{ "0403/astro-ph0403538.txt": { "abstract": "We present the first sub-kpc $(\\sim0.7''\\approx 850$ pc$)$ resolution \\twco(1-0) molecular line observations of the ISM in the host galaxy of the QSO I~Zw~1. The observations were obtained with the BIMA mm-interferometer in its compact A configuration. The BIMA data are complemented by new observations of the \\twco(2-1) and \\thco(1-0) line with IRAM Plateau de Bure mm-interferometer (PdBI) at $0.9''$ and $1.9''$ resolution, respectively. These measurements, which are part of a multi-wavelength study of the host galaxy of I~Zw~1, are aimed at comparing the ISM properties of a QSO host with those of nearby galaxies as well as to obtain constraints on galaxy formation/evolution models. Our images of the \\twco(1-0) line emission show a ring-like structure in the circumnuclear molecular gas distribution with an inner radius of about 1.2~kpc. The presence of such a molecular gas ring was predicted from earlier lower angular resolution PdBI \\twco(1-0) observations. A comparison of the BIMA data with IRAM PdBI \\twco(2-1) observations shows variations in the excitation conditions of the molecular gas in the innermost $1.5''$ comprising the nuclear region of I~Zw~1. The observed properties of the molecular cloud complexes in the disk of the host galaxy suggest that they can be the sites of massive circumnuclear star formation, and show no indications of excitation by the nuclear AGN. This all indicates that the molecular gas in a QSO host galaxy has similar properties to the gas observed in nearby low luminosity AGNs. ", "introduction": "One of the open questions in galaxy evolution is concerned with the formation of active galactic nuclei (AGN) and its relation to star formation (SF) in early type galaxies. The separation of starburst and AGN components in extragalactic objects %and the investigation of the role of star formation % and the properties of the interstellar medium in the -- especially in host galaxies of quasars and QSOs -- is a key problem in the investigation of evolutionary sequences proposed for AGNs \\citep{nor88,san88,rie88,haa03}. Although it is not exactly known how the host galaxy affects the energy release of the QSO, there is statistical evidence for preceding mergers or current interaction of QSO host galaxies with a companion galaxy \\citep{mcl94,lim99}. It is believed that quasar activity is a common, but short lived, phenomenon in galaxy evolution \\citep{mcl99}. Furthermore, host galaxies of QSOs show enhanced SF activities \\citep{cou98}. The transition between Ultra Luminous Infrared Galaxies (ULIRG) and AGN seems to be continuous, all showing signs of enhanced star formation in their nuclear regions \\citep{gen98,mcl99}. %Observations of abundances, excitation and dynamics of the molecular %interstellar medium in the central regions of AGN are essential, since %the interstellar matter is the ''fuel'' for star formation and to a certain %amount also for the central engine. %High resolution molecular line observations allow a detailed study of %the kinematics %of the cold interstellar medium and the derivation of important %parameters %such as gas masses, surface mass densities and star formation %efficiencies. Millimeter molecular line observations are ideal for studying the mechanisms which transport the molecular gas into the AGN. Observations of abundances, excitation and dynamics of the molecular interstellar medium in the central regions of AGN are essential, since the interstellar matter provides the ''fuel'' for star formation as well as the central engine. Consequently, there are large numbers of high resolution observations available for nearby objects. These observations have revealed the presence of circumnuclear starburst rings in a large number of (nearby) active and IR luminous galaxies \\citep [e.g. NGC 1068,] [] {pla91, gen95, hel03}. The extension of high angular resolution observations of molecular gas emission lines to QSO hosts is imperative in order to understand the connection between local active galaxies and high-z QSOs. The results can be used to refine model predictions of the physical conditions in high z- QSOs \\citep[such as used e.g. in] [] {com99}. Due to the limited angular resolution of single dish mm-wavelength telescopes, % molecular line observations, only interferometric observations at these wavelengths allow insight into the morphology and into the kinematics of molecular clouds in the nuclear region of QSO host galaxies with sufficient angular resolution. High resolution molecular line observations allow for detailed kinematic studies of the cold interstellar medium and the derivation of important parameters such as gas masses, surface mass densities and star formation efficiencies. Observations of multiple CO transitions reveal excitation conditions and thus can be used to constrain the physical conditions in an observed region. They can be used to better constrain the contribution of star formation to the total observed infrared luminosities. % Molecular line observations % , which are practically insensitive to the inclination angle of the %galaxy, %directly trace the distribution of the cold interstellar molecular gas %and its kinematics. %The %interpretation of the far infrared emission of QSO in the framework of %star formation is problematic, since, without additional molecular %line observations, it is difficult to derive the amount of thermal %emission which contributes to the total observed infrared %luminosities. Since molecular clouds are the ``fuel'' of star %formation, and the excitation temperature of molecular clouds in SF %regions is elevated with respect to dark molecular clouds, strong %molecular line emission goes along with SF. \\subsection{The nearby QSO I~Zw~1} \\label{IZw1_Intro} The radio-quiet QSO I~Zw~1 is regarded as the closest QSO which can be used for detailed studies of its host galaxy (Tab. \\ref{tab:izw1}). I~Zw~1 has a systemic velocity of 18,290 km s$^{-1}$, which corresponds to a redshift of 0.0611 \\citep{con85}, or a distance of 255 Mpc \\citep[$h_0=0.72$,] [] {spe03}. The nucleus of I~Zw~1 is extremely bright in the optical \\citep[$M_B$ of -23.45 mag,] [] {sch83, bar89}) and also has very bright X-ray emission \\citep{kru90}. I~Zw~1 belongs to the class of infrared luminous galaxies \\citep[$L_{FIR}=10^{11.3}L\\sun$,] [] {haa03}, however its QSO nature %thus can also serve as a template for high-z starburst galaxies. is not apparent in the FIR. There is a strong indication for interaction between I~Zw~1 and a neighboring companion \\citep[e.g.] [] {lim99,schar03}. A direct comparison between the FIR emission and CO molecular line emission in the host galaxy reveals that the FIR continuum emission of I~Zw~1 is predominantly thermal in nature \\citep{bar89}. The FIR to \\twco(1-0) ratio shows a star formation efficiency of about 30 $L_\\sun/M_\\sun$ in the nuclear region \\citep{eck94}. \\citet{schin98} show that the molecular gas mass in the central $3.3''$ (3.7 kpc) is 2/3 of the total observed mass of $9 \\times 10^9 M_\\sun$. A kinematic analysis of their 1.9'' resolution PdBI \\twco(1-0) observations suggests that the central molecular gas is distributed in a ring of $1.5''$ ($\\sim 1.9$ kpc) diameter. However the observations do not spatially resolve this ring. High resolution NIR observations show that the nuclear NIR spectrum is dominated by emission from the AGN, nevertheless about 25\\% of the NIR emission can be attributed to a nuclear starburst \\citep{schin98}, which might be located in the molecular gas ring. The star formation activity in the host galaxy of I~Zw~1 underlines the important role of starbursts in the evolution of QSOs. %in the spatial images, which is consistent with its predicted radius. %of close to 1''. The possible presence of a starburst ring with %sub-kpc radius in the nucleus of I~Zw~1 % and its possible % role for the energy release in the nucleus a the AGN in the %nucleus emphasizes the need for subarcsecond angular resolution %molecular line observations of the QSO host galaxy. %Follow up PdBI in \\twco(2-1) which have an % angular resolution of 0\".9 however show no spatial details, the predicted ring % is not apparent in the \\twco(2-1) transition. Only sub-arcsecond molecular line observations would allow actual imaging of this possible starburst ring. The extension of BIMA's longest baselines up to a maximum of almost 2 km makes the array the millimeter instrument with the highest angular resolution currently available. Here we present \\twco(1-0) $\\sim 0.7''$ angular resolution observations of the QSO I~Zw~1. This corresponds to a spatial resolution of roughly 850~pc. We also present new observations with the IRAM PdBI in \\twco(2-1) and \\thco(1-0) line at $0\".9$ and $2.''1$ resolution, respectively. \\begin{deluxetable}{lcc} %\\tabletypesize{\\scriptsize} \\tablecaption{Adopted Properties of I~Zw~1 (=PG0050+124) \\label{tab:izw1}} \\tablewidth{0pt} \\tablehead{ \\colhead{Property} & \\colhead{Value}} \\startdata Right ascension (J2000.0) & 00$^h$53$^m$34$^s$.9\\\\ Declination (J2000.0) & +12$\\degr 41'36''.2$\\\\ Inclination (deg) & 38\\\\ Position angle (deg) & 135\\\\ Systemic velocity (km s$^{-1}$) & 18,290\\\\ Distance (Mpc)& 255\\\\ $1''$ & 1.24 kpc \\enddata \\tablenotetext{-}{The coordinates refer to the phase center of the millimeter observations. The inclination and position angle are taken from \\citet{schin98}, while the systemic velocity is from \\citet{con85}. } \\end{deluxetable} %PdBI \\thco(1-0) observations with a resolution of $\\sim 2\"$ angular %resolution. %An interpretation of these %data in the framework of multi wavelengths studies to test starburst %models will be presented in another paper. % (Schinnerer et al., in prep.) ", "conclusions": "Our BIMA \\twco(1-0) observations represent the first molecular line observations of the host galaxy of a QSO with sub-kpc resolution. The molecular gas disk of a QSO is for the first time spatially resolved into individual giant molecular cloud complexes. We observe giant molecular clouds with peak molecular hydrogen column densities of $2.4\\times 10^{23}$cm$^{-2}$ at a distance between $1''$ and $3''$ from the nucleus. With such high column densities these clouds could be actively forming stars. The combination of BIMA observations with new and previously published PdBI CO data allows for the first interferometric multi-transition study in the disk of I~Zw~1. The line brightness ratios in the inner $1.5''$ are consistent with moderately dense cold GMCs, and they are not peaked at the center. This strongly suggests that the AGN has no significant effect on the central molecular material. The exact location of the circumnuclear starburst very likely seen in nuclear NIR spectra (Schinnerer et al. 1998) is still not identified. Possible candidates are the resolved \\twco(1-0) molecular clouds seen by BIMA as well as the more thermalized material traced by the higher line ratios inside the inner $1.5''$. However, any of these molecular clouds are likely sites of a massive starburst which contributes to the observed far-infrared luminosity of $L_{FIR}=10^{11.3}L\\sun$. The distribution and properties of the molecular gas in the host galaxy of the nearby QSO I~Zw~1 are quite similar to what is observed in nearby low luminosity AGNs \\citep[e.g.] [] {pag01, gar03}. This suggests that the ISM in QSO at high redshift might be similar to nearby low luminosity AGN as well. Four additional attempts to re-observe I~Zw~1 in \\twco(1-0) with BIMA in the A configuration all failed to provide useful data, due to insufficient phase coherence of the observations. This demonstrates that the observations presented here are on the limit of what can be done without active phase correction systems, at least from a site such as Hat Creek where the BIMA array is situated. %The molecular line emission to the south-east of the galaxy center is stronger than the emission to the north-west. The reason for this %is rather that the emission to the north is resolved out by the long BIMA baselines, %whereas it is not by the lower resolution PdBI observations. %-- Gas masses: Schinnerer et al.: %Schinnerer et al. (1998) observe a \\twco(1-0) line flux of 56 mJy~beam$^{-1}$ %from the %innermost 3.''3 which is the emission from the nuclear region. %They find that the $N_{H_2}/I_{CO}$ %conversion factor in I~Zw~1 is close to $2\\times10^{20} $cm$^{-2}$ K$^{-1}$ km$^{-1}$ s %which is also found for molecular gas in our galaxy and and many %nearby external galaxies. %The derived molecular nucleus hydrogen column density in the inner %3\".3 is $4.2\\times10^{22}$ cm$^{-2}$, the molecular gas mass is $7.5\\times10^9 %M_\\sun$. % This is 2/3 of the total measured flux. The corrsponding mass is $9 % \\times 10^9 M_\\sun$. %--The kinemattic gas masses derived from the rotation curve indicate a %CO/mass ratio which %is in good agreement with Galactic conversion factors. Our data ..... %-- a comparison with the high resolution \\twco(2-1) data of Eckart %et al. show that %the molecular inside the 1 kpc molecular ring has %higher temperatures/densities than the molecular gas in the ring. This is consistent %with the finding, that %-- topology: %Our integrated line maps with different synthesized beam weights. %There is strong clumping of the molecular material, i.e. a majority %(... Jy compared to ... Jy in a 2'' beam) are detected at a %resolution of ...) %emission from small scale structures (structures with high spatial %frequencies) %--Fig. .. shows an overlay of the integrated CO 1-0 molecular line %emission and an optical image obtained with ...... %-- We compare our fluxes for different synthesized beam sizes to the %Schinnerer % et al. fluxes: %The corresponding brightness temperature of $T_B = 6$ K indicates a %large beam filling factor and thermalization of a significant amount %of gas within an area corresponding to roughly 500 pc across. % kinematische Masse: v^2=GM/R %-- discuss numbers %convert table from power point talk) %-- simplest radiative transfer? Wolfire model calcs? %-- compare with other published numbers %-- A look at channel maps reveals: channel maps are consistent with %orbital % motion. %-- velocity diagrams %-- masked data pos-vel maps %-- discuss spectrum %We find a double looped spectrum which is consistent with earlier %published %spectra. %A comparison of our observed spectrum with the lower resolution %spectrum from %single dish observations ..." }, "0403/astro-ph0403697_arXiv.txt": { "abstract": "Recently, there has been a great deal of interest in understanding the reionization of hydrogen in the intergalactic medium (IGM). One of the major outstanding questions is how this event proceeds on large scales. Motivated by numerical simulations, we develop a model for the growth of \\ion{H}{2} regions during the reionization era. We associate ionized regions with large-scale density fluctuations and use the excursion set formalism to model the resulting size distribution. We then consider ways in which to characterize the morphology of ionized regions. We show how to construct the power spectrum of fluctuations in the neutral hydrogen field. The power spectrum contains definite features from the \\ion{H}{2} regions which should be observable with the next generation of low-frequency radio telescopes through surveys of redshifted 21 cm emission from the reionization era. Finally, we also consider statistical descriptions beyond the power spectrum and show that our model of reionization qualitatively changes the distribution of neutral gas in the IGM. ", "introduction": "\\label{intro} The reionization of the intergalactic medium (IGM) is one of the landmark events in early structure formation. It marks the epoch at which radiative feedback from luminous objects impacted the farthest reaches of the Universe -- the point at which structure formation affected every baryon in the Universe, albeit indirectly. The timing, duration, and character of this event contain an enormous amount of information about the first cosmic structures and also have important implications for later generations of baryonic objects. For these reasons, a great deal of attention -- both observational and theoretical -- has recently been focused on this process. Most significantly, reionization is an important signpost that connects several disparate observations. Currently the data provide tantalizing hints about the ionization history of the Universe but few definitive answers. Observations of Ly$\\alpha$ absorption in the spectra of high-redshift quasars indicate that reionization ends at $z \\sim 6.5$ \\citep{becker,fan,white03,wyithe04-prox}, although this interpretation is controversial \\citep{songaila04}. The main difficulty with these measurements is that the Ly$\\alpha$ optical depth is extremely large in a fully neutral medium \\citep{gp}, making it difficult to place strong constraints when the neutral fraction exceeds $\\sim 10^{-3}$. On the other hand, measurements of the cosmic microwave background (CMB) imply a high optical depth to electron scattering, apparently requiring reionization to begin at $z \\ga 14$ \\citep{kogut03,spergel03}. Unfortunately, the CMB data provide only an integral constraint on the ionization history. Taken together, these observations rule out simple pictures of fast reionization (e.g., \\citealt{barkana01} and references therein), but it is not yet clear what they do imply about early generations of luminous sources \\citep{sokasian03a,sokasian03b,wyithe03,cen03,haiman03,onken03,fukugita03}. At the same time, our theoretical understanding of how reionization proceeds, given some source population, has been advancing rapidly. Most models of reionization are based on the growth of \\ion{H}{2} regions around individual galaxies \\citep{arons72,barkana01}. These models use semi-analytic techniques to compute the evolution of global quantities like the total ionized fraction. However, they are unable to describe the morphology of reionization, an issue that is both theoretically interesting and observationally accessible. Morphology is a difficult aspect to address because it depends on such factors as the locations of individual sources, the clumpiness of the IGM, the underlying density distribution (the cosmic web), recombinations, and radiative transfer effects. Some semi-analytic models have been developed to address these issues. The most popular assumes that reionization is controlled by recombinations and proceeds from low to high density regions \\citep{miralda00}, but these models are approximate at best. Fortunately, it has recently become possible to incorporate radiative transfer into numerical simulations of the reionization era \\citep{gnedin00,sokasian03a,sokasian03b,ciardi03-sim}, at least as a post-processing step. Because simulations can include all of the above processes (with the possible exception of clumpiness, which still requires extremely high mass resolution), they give a more nuanced view of the reionization process. One of the chief lessons of the simulations is that reionization is significantly more inhomogeneous than expected. The classical picture of a large number of small \\ion{H}{2} regions surrounding individual galaxies does not match the simulations well; instead, a relatively small number of large ionized regions appear around clusters of sources (see, for example, Figure 6 of \\citealt{sokasian03a}). Moreover, in the simulations reionization proceeds from high to low density regions, implying that recombinations play only a secondary role in determining the morphology of reionization; instead, large-scale bias plays a dominant role. This picture suggests a new approach to analytic models of reionization, one that takes into account the large-scale density fluctuations that are ultimately responsible for this ionization pattern. We describe such a model in \\S \\ref{bubbles}. We derive the size distribution of \\ion{H}{2} regions in a way analogous to the \\citet{press} halo mass function and show that it has the qualitative features seen in the simulations. Observing the morphology of \\ion{H}{2} regions requires new observational techniques as well as robust ways to characterize the data. Among the most exciting approaches to study the reionization process are surveys of 21 cm emission from neutral hydrogen at high redshifts \\citep{field58,scott,mmr,zald04}. The idea behind these observations is to map the fluctuating pattern of emission (or absorption) from neutral hydrogen in the Universe over a range of frequencies. This yields a measurement of fluctuations due to both the density field and the \\ion{H}{2} regions. Because 21 cm emission comes from a single spectral line, such observations allow a three-dimensional reconstruction of the evolution of neutral hydrogen in the IGM and, owing to the design of low-frequency radio telescopes, can probe the large cosmological volumes needed to study the IGM. Despite numerous technical challenges, instruments able to make the necessary observations will be built in the coming years. These include the Primeval Structure Telescope (\\emph{PAST}),\\footnote{ See http://astrophysics.phys.cmu.edu/$\\sim$jbp for details on PAST.} the Low Frequency Array (\\emph{LOFAR}),\\footnote{ See http://www.lofar.org for details on LOFAR.} and the Square Kilometer Array (\\emph{SKA}).\\footnote{ See http://www.skatelescope.org for details on the SKA.} The major obstacles to high-redshift 21 cm observations are the many bright foreground sources, which include Galactic synchrotron emission, free-free emission from galaxies \\citep{oh03}, faint radio-loud quasars \\citep{dimatteo02}, and synchrotron emission from low-redshift galaxy clusters \\citep{dimatteo04}. Fortunately, all of these foregrounds have smooth continuum spectra. Zaldarriaga et al. (2004; hereafter ZFH04) showed that the foregrounds can be removed to high precision because the 21 cm signal itself is uncorrelated over relatively small frequency ranges. ZFH04 also showed how to compute the angular power spectrum of the 21 cm sky as a function of frequency, given some model of reionization. This is the simplest statistical measure of the morphology of \\ion{H}{2} regions, and they showed that it is quite powerful in distinguishing different stages of reionization (at least in a simple toy model). Of course, predictions for the 21 cm signal, or any other measurement of reionization, rely on an accurate model for reionization. The initial conditions are straightforward: when the neutral fraction $\\bxh \\approx 1$, the power spectrum follows that of the density \\citep{tozzi}. This breaks down when the \\ion{H}{2} regions appear, and more sophisticated approaches become necessary. There have been several recent attempts to predict the signal using numerical simulations \\citep{ciardi03,furl-21cmsim,gnedin03}. However, the boxes in these simulations have sizes $\\la 20 \\Mpc$, subtending only a few arcminutes on the sky. Because the angular resolution of low frequency radio telescopes at the required sensitivity is also a few arcminutes (ZFH04), such predictions require extrapolations to larger scales, which may be dangerous given the already large sizes of ionized regions. With our analytic model for the size distribution of these bubbles, we are able to make the first detailed statistical characterizations of the fluctuating ionization pattern on the large scales most relevant to observations. We begin with the power spectrum as the simplest description. In \\S \\ref{ps}, we show how to compute the power spectrum for an arbitrary size distribution, and in \\S \\ref{res} we apply that formalism to our model of \\ion{H}{2} regions. The power spectrum is only one way to describe the 21 cm field. It is an excellent approximation if fluctuations in the matter density dominate the signal, because the density field is neary gaussian on the large scales of interest here. Unfortunately, once the \\ion{H}{2} regions dominate the power spectrum, the probability distribution is no longer gaussian \\citep{morales03}. There have, however, been no attempts to describe the distribution. Using our model for reionization, we discuss some of the relevant features in \\S \\ref{nongauss}. We show that the reionization model qualitatively changes the character of fluctuations in the neutral gas density. This paper is primarily concerned with developing a general and useful model for the morphology of reionization and with the crucial features of that model. We will focus here on the physics of reionization rather than on their observable consequences; we consider some of these in a companion paper about the 21 cm signal expected from high redshifts \\citep{furl04b}. In our numerical calculations, we assume a cosmology with $\\Omega_m=0.3$, $\\Omega_\\Lambda=0.7$, $\\Omega_b=0.046$, $H=100 h \\hunits$ (with $h=0.7$), $n=1$, and $\\sigma_8=0.9$, consistent with the most recent measurements \\citep{spergel03}. ", "conclusions": "\\label{disc} In this paper, we have described a new analytic model for the size distribution of \\ion{H}{2} regions during reionization. While the overlap process is usually described in terms of Str{\\\" o}mgren spheres around individual galaxies, recent cosmological simulations of reionization have demonstrated that the ionized regions are much larger than naively expected, even early in the reionization process. Here we have shown that the size distribution can be understood in terms of large-scale features in the density field, and we have constructed the size distribution with an approach analogous to the standard derivation of the \\citet{press} mass function for collapsed halos. The model has only two input parameters (if the cosmology is fixed): the ionizing efficiency of collapsed objects and the minimum mass halo that can host a luminous object. Interestingly, at a fixed neutral fraction the characteristic size of the bubbles is fairly insensitive to these input parameters, suggesting that the morphology of reionization is close to invariant (at least for simple, single reionization episodes). While we make several simplifying approximations in the model (see the discussion in \\S \\ref{bubbles}), it provides a self-consistent approach that reproduces the qualitative features of simulations. In the future, the model must be quantitatively compared to simulations; however, an accurate comparison requires simulations with large ($\\sim 100^3 \\Mpc^3$) volumes because of the effects of the large-scale density field (see Figure \\ref{fig:dndr-comp}, as well as \\citealt{barkana03}). The morphology of reionization has implications for a number of observables. Most important, regardless of the technique, we predict large \\ion{H}{2} regions that should be feasible to detect, whether through quasar absorption spectra \\citep{miralda00,barkana02}, Ly$\\alpha$ lines at extremely high-redshifts \\citep{pello04,loeb04,ricotti04,gnedin04,cen04,bart04}, or 21 cm tomography, For example, the noise in a 21 cm map is proportional to the square root of the bandwidth and, more important, to the square of the pixel size. Our model predicts substantial contrast on relatively large scales of several arcminutes, which should make detections easier. Note, however, that in many next-generation experiments like LOFAR the ``bandwidth\" can be chosen \\emph{after} the observations are complete, so the expected scale of the bubbles need not determine the experimental design \\citep{morales03}. Even if high signal-to-noise detections of individual \\ion{H}{2} regions are not available, we have shown that statistical measurements of the size distribution can still strongly constrain reionization. We first constructed the power spectrum of fluctuations in the neutral density, including both density fluctuations and the ionized regions. We found that the ionized regions imprint clear features on the power spectrum and amplify the power by a factor of several during the middle and late stages of reionization. Most important, the power spectrum evolves throughout reionization, allowing us to map the time history of reionization. For 21 cm observations, the large-scale \\ion{H}{2} regions predicted by our model put the features at $l \\la 10^4$ (or $\\theta \\ga 2\\arcmin$). This matches well to the scales able to be probed by upcoming experiments like PAST, LOFAR, and SKA (ZFH04). If, on the other hand, reionization occurred through the overlap of \\ion{H}{2} regions around individual galaxies, the features in the power spectrum would appear at much smaller scales, perhaps beyond the reach of these instruments. We have also shown explicitly that the ionized bubbles induce qualitative changes to the initially gaussian neutral density distribution. This suggests that statistical measurements beyond the power spectrum can offer probes of the physics of reionization that may be even more powerful. Our results also have important implications for other measurements. For example, one of the most successful strategies for targeting high-redshift galaxies is by searching for strong Ly$\\alpha$ emitters. If the galaxy is embedded in a mostly neutral medium, Ly$\\alpha$ absorption from the IGM can have a substantial effect on the line profile \\citep{haiman02,santosm03,loeb04}. In our model, we expect this absorption to be less significant, because most galaxies are embedded in \\ion{H}{2} regions with large sizes, even relatively early in the reionization process. Finally, in this paper we have considered only the simplest reionization histories with a single type of source. Many models for reconciling the quasar and CMB data on reionization require multiple, distinct generations of sources that cause ``stalling\" or even ``double\" reionization \\citep{wyithe03,cen03,haiman03,sokasian03b}. Such histories will of course change the morphology of reionization and hence modify the 21 cm signal. Moreover, alternative models of reionization in which (for example) voids are ionized first yield different sets of signatures. In \\citet{furl04b}, we use the formalism developed here to examine how well 21 cm tomography can distinguish these histories and models." }, "0403/astro-ph0403374_arXiv.txt": { "abstract": "The mergers of black hole-neutron star binaries are calcuated using a pseudo-general relativistic potential that incorporates ${\\mathcal O}(v^2/c^2)^3$ post-Newtonian corrections. Both normal matter neutron stars and self-bound strange quark matter stars are considered as black hole partners. As long as the neutron stars are not too massive relative to the black hole mass, orbital decay terminates in stable mass transfer rather than an actual merger. For a normal neutron star, mass transfer results in a widening of the orbit but the stable transfer ends before the minimum neutron star mass is reached. For a strange star, mass transfer does not result in an appreciable enlargement of the orbital separation, and the stable transfer continues until the strange star essentially disappears. These differences might be observable through their respective gravitational wave signatures. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403468_arXiv.txt": { "abstract": "Although thermal disk emission is suppressed or absent in the hard state of X-ray binaries, the presence of a cold, thin disk can be inferred from signatures of reprocessing in the $\\sim2-50$ keV band. The strength of this signature is dependent on the source spectrum and flux impinging on the disk surface, and is thus very sensitive to the system geometry. The general weakness of this feature in the hard state has been attributed to either a truncation of the thin disk, large ionization, or beaming of the corona region away from the disk with $\\beta\\sim0.3$. This latter velocity is comparable to jet nozzle velocities, so we explore whether a jet can account for the observed reflection fractions. It has been suggested that jets may contribute to the high-energy spectra of X-ray binaries, via either synchrotron from around $100-1000$ $r_{\\rm g}$ along the jet axis or from inverse Compton (synchrotron self-Compton and/or external Compton) from near the base. Here we calculate the reflection fraction from jet models wherein either synchrotron or Compton processes dominate the emission. Using as a guide a data set for GX~339$-$4, where the reflection fraction previously has been estimated as $\\sim10\\%$, we study the results for a jet model. We find that the synchrotron case gives $< 2\\%$ reflection, while a model with predominantly synchrotron self-Compton in the base gives $\\sim 10-18\\%$. This shows for the first time that an X-ray binary jet is capable of significant reflection fractions, and that extreme values of the reflection may be used as a way of discerning the dominant contributions to the X-ray spectrum. ", "introduction": "X-ray binaries (XRBs) have been observed in several distinct states, which are characterized by the relative strength of their soft and hard X-ray emission components, as well as by their variability properties \\citep[see, e.g.,][]{McClintockRemillard2003}. In the ``standard'' models \\citep[see][and references therein]{ReynoldsNowak2003} the soft component is well-explained with thermal emission from a standard thin disk \\citep{ShakuraSunyaev1973}, while the hard power-law component is generally attributed to inverse Compton (IC) scattering processes. The various models currently in existence often have quite different seed photons and system geometries, yet predict similar results for the broad continuum emission \\citep[see][]{NowakWilmsDove2002,Markoffetal2003}. In order to discern between the models, therefore, one has to look at finer details which are dependent upon specific elements of the geometry. In the hard state, the power-law component dominates the thermal disk emission over most of the X-ray range. The presence of a cold, thin disk can still be inferred via detection of a soft component in the $\\approx$0.3--1\\,keV band, and via spectral components in the $\\approx$2--50\\,keV band suggesting that a fraction of the hard X-rays is reprocessed or reflected from an optically thick surface. The reflection component is characterized by a flattening of the power-law above $\\sim10$\\,keV \\citep[e.g.,][]{Poundsetal1990,GeorgeFabian1991}, as well as by spectral features such as an Fe K$\\alpha$ fluorescent line and an Fe edge \\citep[see][for a review]{ReynoldsNowak2003}. The strength of these components is directly related to the spectrum and flux hitting the disk, and is therefore sensitive to assumptions about the system geometry. For example, in models where the hard X-rays are due to IC in a hot coronal plasma completely ``sandwiching'' the disk, the reflection is easily too high for typical X-ray binary spectra (and the self-consistently derived coronal temperatures are too low; e.g., \\citealt{stern:95a,dove:97a}). Thus modifications have been proposed, such as a recessed thin disk \\citep{dove:97b,poutanen:97b}, beaming away from the disk \\citep{reynolds:97c,Beloborodov1999}, or large amounts of ionization of the disk \\citep{ross:99a,nayakshin:01a,done:01c}. In addition to the standard corona models, \\citet{MarkoffFalckeFender2001} proposed that the entire broadband spectrum of hard state XRBs could instead result from synchrotron emission at the beginning of an acceleration region. While controversial, this model succeeds at explaining the tight correlation of radio and X-ray emission seen in several sources \\cite[e.g., GX~339$-$4,][]{Corbeletal2000,Corbeletal2003,Markoffetal2003}, which, in fact, may be a universal correlation in XRB hard states \\citep{GalloFenderPooley2003}. It is also the first model to provide a link between the inferred presence of a hot, magnetized electron plasma near the inner regions of the central engine to the hot, magnetized plasma that we know exists via imaged radio jets \\citep[e.g.,][]{Stirlingetal2001}. A shortcoming of these models, however, is that they have not yet attempted to explain detailed X-ray spectral features such as reflection and iron lines. In this Letter we calculate the expected reflection fraction from several jet models with parameters that provide good descriptions of broad band features (e.g., overall luminosity, flat radio spectrum, X-ray spectral slope and cutoff) for `typical' XRB data sets, such as those shown here for the Galactic source GX~339$-$4. We then discuss the ensuing constraints on jet models. ", "conclusions": "As a first step towards judging the magnitude of the reflection fraction, we can calculate $R(\\mu)$ as a function of energy by substituting $dP/d\\Omega d\\nu$ into eq.~\\ref{eq:def}. The synchrotron-dominated cases show the smallest overall reflection. The ratio is only $\\approx$1--2\\% in between energies of 1--100\\,keV ($\\beta_{\\rm s}=0.3, ~0.4$). The ratio peaks at 0.3\\,keV with value 6\\% and 8\\% for $\\beta_{\\rm s}=0.3$ and $\\beta_{\\rm s} = 0.4$, respectively. The SSC-dominated jet shows significantly more reflection. For $\\beta_{\\rm s}=0.3$ and $\\beta_{\\rm s}=0.4$, the ratio is $>10\\%$ in between 0.5--23\\,keV. The former peaks at 18\\% at 6\\,keV, while the latter peaks at 17\\% at 5\\,keV. Such values of `reflection fraction' are comparable to the observed range for GX~339$-$4, and in fact are larger than fitted for this particular observation \\citep{NowakWilmsDove2002}. \\clearpage \\begin{figure*} \\epsscale{1} \\plottwo{f4a.eps}{f4b.eps} \\caption{Total model spectrum (direct plus reflected - solid line), summed spectrum incident upon the disk (dashed line), and reflected spectrum (lower solid line). {\\it Left:} SSC-dominated jet with $\\beta_{\\rm s} = 0.4$. {\\it Right:} Synchrotron-dominated jet with $\\beta_{\\rm s} = 0.4$. \\label{fig:ref}} \\end{figure*} \\clearpage As described above, we have also directly calculated the reflection spectrum expected from our models. We show spectra for the SSC-dominated jet and the synchrotron-dominated jet, both with $\\beta_{\\rm s} = 0.4$, in Fig.~\\ref{fig:ref}. Qualitative differences are immediately apparent. The SSC-dominated jet clearly has an overall brighter reflection spectrum, and effects of relativistic smearing are readily visible. These effects are to be expected given both the lower velocities of the SSC-dominated region ($\\beta \\sim 0.3$) as well as the fact that this region is closer to the central black hole. We have chosen a slightly face-on orientation, $\\mu=0.77$, since there has been some suggestion that this is indeed the case for the GX~339$-$4 system \\citep{Wuetal2001}. Most BHC systems should have a higher inclination, which, given eq.~\\ref{eq:def}, allows for even greater reflection fractions. If we freeze all of our model parameters but instead choose $\\mu=0.5$, reflection fractions, as defined by the ratio of the disk-incident spectrum to the directly viewed spectrum, increase. Over the 1--20\\,keV interval, this ratio is everywhere $> 26\\%$ for the SSC-dominated jet with $\\beta_{\\rm s} = 0.3$, and $>7\\%$ for the synchrotron-dominated jet with $\\beta_{\\rm s} = 0.4$. The former peaks at 38\\% at 2\\,keV and, again, exhibits clear relativistic broadening of the reflection features. While still preliminary, this is the first time that the reflection of jet emission off an accretion disk has been calculated for an X-ray binary system. These results show that jets are indeed capable of producing the reflection fraction inferred from the X-ray data, providing further support for a connection between the base of the jet and the corona. We have highlighted two extreme cases which might be applicable to different physical situations. Systems that exhibit very low reflection ($\\lesssim 5\\%$), with sharp (i.e., non-relativistic) features in any reflected spectrum \\citep[e.g., XTE J1118+480,][]{Milleretal2001}, could be synchrotron-dominated, and clearly rule out SSC-dominated jets. Systems with significantly larger reflection fractions ($\\gtrsim 15\\%$) cannot be synchrotron-dominated, especially if they exhibit features which are unambiguously relativistically smeared. However even in the Compton-dominated regime, as shown in Fig.~\\ref{gx339}, synchrotron radiation can contribute $\\gtrsim 10\\%$ of the flux which, as we discuss further below, will greatly effect fits to data with corona plus reflection models. For intermediate values, or values $\\gtrsim 30\\%$, other factors need to be considered. For instance, these results assume that the disk is perfectly flat, and that the jet is always perpendicular to the disk. Realistically, disks are expected to flared or warped \\citep[e.g.,][]{Dubusetal1999}, and several systems also show evidence for misalignment between the jets and outer disks \\citep{Maccarone2002}. Both of these effects will serve to increase the reflection fraction from the jet, particularly for the synchrotron component. Therefore, we treat these numbers as lower limits. We would like to emphasize, however, that a significant X-ray contribution from jet synchrotron emission can greatly alter how one even defines `reflection fraction' based upon a presumed single-component, underlying continuum. As shown in Fig.~\\ref{cygx1}, jet radiation alone can provide a good description of the high energy cutoff region. One can readily imagine a model wherein the soft X-ray region is dominated by SSC emission (as in Fig~\\ref{fig:ref}b) and/or Comptonization of external (disk) photons, each with a large covering factor of the disk (i.e., essentially unity reflection fraction, for that component alone). The hard X-ray radiation could then be dominated by synchrotron radiation with inherently low reflection fraction. The net spectrum would have an intermediate fitted value of `reflection fraction' that does not have a `geometric interpretation' entirely appropriate for either the soft or hard emission components. If the broad-band X-ray continua of hard state BHC are in fact comprised of such multiple components, as in some of the jet models presented here, then this calls into question current interpretations of `reflection fractions' based upon single component fits. Of course, in order to determine whether such multiple spectral components are indeed present in the observations, actual fitting of the combined direct plus reflection spectrum needs to be performed. The calculations and models presented here provide vital clues as to how much each process can contribute for this next step. This work has shown that this type of analysis may hold the key to disentangling the emission processes relevant from the accretion inflow and outflow, and place limits on the synchrotron vs. Compton contributions to the hard state spectrum." }, "0403/astro-ph0403142_arXiv.txt": { "abstract": "{ We present the initial results from our \\xmm\\ program aimed at searching for X-ray activity cycles in solar-type stars. \\hd\\ is a G2-type star (somewhat more evolved than the Sun, and with a less massive companion) with a pronounced 8.2 yr chromospheric cycle, as evident from from the Mt.\\ Wilson program data. We present here the results from the initial 2.5 years of \\xmm\\ observations, showing that large amplitude (a factor of $\\simeq 10$) modulation is present in the X-ray luminosity, with a clearly defined maximum in mid 2002 and a steady decrease since then. The maximum of the chromospheric cycle took place in 2001; if the observed X-ray variability is the initial part of an X-ray cycle, this could imply a phase shift between chromospheric and coronal activity, although the current descent into chromospheric cycle minimum is well reflected into the star's X-ray luminosity. The observations presented here provide clear evidence for the presence of large amplitude X-ray variability coherent with the activity cycle in the chromosphere in a star other than the Sun. ", "introduction": "\\label{sec:intro} The 11 year cycle is perhaps the oldest known manifestation of the Sun's magnetic activity, having first been noted by Schwabe in 1843 as a periodic modulation of the number of sunspots. Subsequently, most activity indicators have been observed to follow a similar cyclical variation, with the amplitude of the modulation dependent on the indicator used. On cool stars other than the Sun, the detection of cycles had to wait for the foresight of O. Wilson, who started a long-term monitoring of a Ca\\,{\\sc ii} H\\&K activity indicator (the ``$S$ index'') in a significant number of stars, using the Mt. Wilson 100 inch telescope. An analysis of the vast amount of data from the Mt. Wilson program, now covering nearly 40 years of observations (\\citealp{bds+95}), shows that solar-like cycles are present in many stars, although some stars show no detected variability (perhaps being in a ``Maunder-like'' state) while others show significant non-periodic variability. In the Sun, the amplitude of the cyclical S index modulation is a factor of about 2, while the amplitude in X-rays is much stronger, i.e.\\ a factor of 100 in the Yohkoh 0.73--2.5 keV band. Yet evidence of cyclical variability in X-rays in stars other than the Sun has only become available very recently. In fact (with the recent exception of 61~Cyg), the X-ray observations of the few stars for which sufficient data exist suggest that their X-ray luminosity is relatively stable over long-term intervals, as discussed e.g. by \\cite*{ste98b}. Observations of homogeneous samples of active stars, e.g.\\ of the Hyades, show little variation in \\lx\\ across the $\\simeq 10$-year separation between the \\emph{Einstein} and ROSAT PSPC surveys, with 90\\% of the stars showing less than a factor of 2 variability. In these stellar samples, however, the median activity level is much higher than the Sun's, by 2 or more dex: stars at this activity level typically do not show cycles in Ca\\,{\\sc ii}, instead varying irregularly, so that perhaps the lack of X-ray modulation is not surprising. This result was however confirmed by \\cite*{ls2004} on a volume-limited sample of solar-type stars using ROSAT All-Sky Survey (RASS) and pointed observations. Weak statistical evidence for the presence of solar-like cycles in less active stars was derived by \\cite*{hss96} using the RASS observations of the stars in the Mount Wilson program, looking at the deviations from a ``mean'' X-ray luminosity for each star as a function of the Ca\\,{\\sc ii} cycle phase. In their analysis of X-ray variability properties of solar mass stars \\cite*{mmp+2002} found that, in relatively quiet stars, amplitude variations increase with time scales, and interpreted this as an indication of the presence of solar-like cycles in stars with X-ray activity of the same order of that of the Sun. The comparison between the Sun and nearby stars is consistent with a fraction of moderately active stars ($\\overline{L}_{\\rm X} < 10^{28}$ erg/sec) having X-ray variability similar to the Sun, while more active stars lack solar-like cyclic coronal activity (\\citealp{mm2003}). More recently, \\cite{hsb+2003} presented evidence of X-ray luminosity modulation in 61 Cyg A and B (K5V and K7V) over 4.5 years, well correlated with their $S$ index, strongly suggestive of an activity cycle in X-rays. The observed modulation amplitude (using ROSAT HRI data, which prevented a study of the spectral evolution) is a factor of 2.5. In this paper we present the results of the first 2.5 yr of the \\hd\\ \\xmm\\ monitoring program, showing for the first time in a star other than the Sun strong evidence for long-term coherent X-ray variability with an amplitude a factor of about 10, consistent with the cyclic amplitude modulation in the Sun. ", "conclusions": "\\label{sec:concl} We have presented clear evidence of long-term variations of the X-ray luminosity in \\hd, a solar-type star with a well defined cycle in its chromospheric activity. The variations thus far determined have an amplitude of a factor of 10 (over the 2.5 years covered by the observations), comparable to the variations seen in the X-ray luminosity of the Sun during the solar cycle. These variations suggest the beginning of a cycle; while the observed X-ray maximum appears somewhat offset (by about 1 year) from the chromospheric one, the current descending phase of the chromospheric cycle is well reflected in \\hd's decreasing X-ray luminosity over the last two years. \\hd\\ is the subject of a long-term monitoring program performed with the \\xmm\\ observatory (which also includes $\\alpha$ Cen and 61~Cyg), of which we are presenting here the first (still perforce preliminary) results. A further two years of observations (at six months cadence) are already planned on the same target, and we will re-propose the target for every AO, to ensure the continuous monitoring. While the nature of this program is such that a few more years will be necessary before a detailed analysis can be performed, the initial results presented here show that our initial choice of targets was a good one; we now can state with some confidence that the Sun is not the only solar-type star for which large amplitude X-ray variability coherent with the activity cycle in the chromosphere is present. The continuation of our monitoring program on \\hd\\ will allow to compare the characteristics of its coronal cycles with the solar one." }, "0403/astro-ph0403232_arXiv.txt": { "abstract": "Long-slit stellar kinematic observations were obtained along the major-axis of $30$ edge-on spiral galaxies, $24$ with a boxy or peanut-shaped (B/PS) bulge and $6$ with other bulge types for comparison. B/PS bulges are identified in at least $45\\%$ of highly inclined systems and a growing body of theoretical and observational work suggests that they are the edge-on projection of thickened bars. Profiles of the mean stellar velocity $V$, the velocity dispersion $\\sigma$, as well as the asymmetric ($h_3$) and symmetric ($h_4$) deviations from a pure Gaussian are presented for all objects. Comparing those profiles with stellar kinematic bar diagnostics developed from N-body simulations, we find bar signatures in $24$ of our sample galaxies ($80\\%$). Galaxies with a B/PS bulge typically show a double-hump rotation curve with an intermediate dip or plateau. They also frequently show a rather flat central velocity dispersion profile accompanied by a secondary peak or plateau, and numerous galaxies have a local central $\\sigma$ minimum ($\\gtrsim40\\%$). The $h_3$ profiles display up to three slope reversals. Most importantly, $h_3$ is normally correlated with $V$ over the presumed bar length, contrary to expectations from axisymmetric disks. These characteristic bar signatures strengthen the case for a close relationship between B/PS bulges and bars and leave little room for other explanations of the bulges' shape. We also find that $h_3$ is anti-correlated with $V$ in the very center of most galaxies ($\\gtrsim60\\%$), indicating that these objects additionally harbor cold and dense decoupled (quasi-)axisymmetric central stellar disks, which may be related to the central light peaks. These central disks coincide with previously identified star-forming ionized-gas disks (nuclear spirals) in gas-rich systems, and we argue that they formed out of gas accumulated by the bar at its center through inflow. As suggested by N-body models, the asymmetry of the velocity profile ($h_3$) appears to be a reliable tracer of asymmetries in disks, allowing us to discriminate between axisymmetric and barred disks seen in projection. B/PS bulges (and thus a large fraction of all bulges) appear to be made-up mostly of disk material which has acquired a large vertical extent through bar-driven vertical instabilities. Their formation is thus probably dominated by secular evolution processes rather than merging. ", "introduction": "} The number of edge-on disk galaxies with a boxy or peanut-shaped (B/PS) bulge is known to be significant. In fact, the fraction of these objects has continuously increased as sample selection, data, and classification criteria improved \\citep*[see][]{j86,sa87,s87,db88,ldp00a,ldp00b}. Recently, \\citet{ldp00a} classified $\\approx1350$ edge-on disk galaxies and found $45\\%$ of B/PS bulges, with a fraction higher than $40\\%$ for all morphological types from S0 to Sd. This was the first study to report such a large incidence of B/PS bulges in very late type systems, but it is also the largest and most homogeneous. B/PS bulges are therefore common, but they are only seen in highly inclined systems ($i\\gtrsim75^{\\circ}$; e.g.\\ \\citealt*{sdb90}), indicating that their shape is mainly related to the vertical distribution of light. Several scenarios have been suggested to explain the structure and formation of B/PS bulges. They can be divided into two main categories: axisymmetric accretion and non-axisymmetric bar buckling models. \\citet{bp85} and \\citeauthor{r88} (\\citeyear{r88}; but see also \\citealt{man85}) showed that it is possible to form axisymmetric cylindrically rotating B/PS bulges from the accretion of one or (more likely) many satellite galaxies onto a larger galaxy host. As pointed out by \\citet{bp85}, however, the particular alignment required puts too many constraints on the orbits, angular momenta, and relative masses of the galaxies, and this process appears unlikely in view of the high incidence of B/PS bulges. Moreover, satellite galaxies are not preferentially observed around galaxies with a B/PS bulge \\citep{s87}. Using N-body simulations, \\citet{cs81} first suggested that B/PS bulges could be due to the thickening of bars in the disks of spiral galaxies. They showed that soon after a bar develops through the usual bar instability, it buckles, thickens, and settles with a larger velocity dispersion and thickness. When highly inclined, the bar then appears peanut-shaped when seen side-on (i.e.\\ perpendicular to the bar), boxy-shaped when seen at intermediate angles, and almost round when seen end-on (i.e.\\ parallel to the bar). It is also largely cylindrically rotating. \\citeauthor{cs81}'s (\\citeyear{cs81}) results were later confirmed and improved upon by many studies \\citep[e.g.][]{cdfp90,rsjk91,am02} and they are now universally accepted. The three-dimensional (3D) orbital structure of these bars has just been investigated in a series of papers (\\citealt*{spa02a,spa02b,psa02,psa03}; but see also \\citealt{p84,pf91}), allowing a more detailed comparison of models with high spatial resolution images. There are several observational studies supporting the connection between B/PS bulges and bars. Most important for this work, \\citet{km95} suggested that bars could be detected in edge-on galaxies from their (projected) kinematic signature. In their pioneering study of two spirals with a peanut-shaped bulge (\\objectname{NGC~5746} and \\objectname{NGC~5965}), \\citet{km95} found a characteristic ``figure-of-eight'' bar signature in their position-velocity diagrams (PVDs; the projected density of material as a function of line-of-sight velocity and projected position). Although the interpretation of this signature was ultimately flawed, hydrodynamical simulations by \\citet{ab99} confirmed its relationship to bars and showed how it can be used to identify and characterize edge-on bars from gas kinematics. In a study of emission-line spectra for $10$ edge-on spirals with various types of bulges, \\citet{mk99} later confirmed the link between B/PS bulges and bars. \\citeauthor{bf99} (\\citeyear{bf99}; see also \\citealt{b98}) also presented ionized-gas PVDs for $23$ edge-on galaxies, $17$ with a B/PS bulge and $6$ with other types of bulges (mostly spheroidal). Using \\citeauthor{ab99}'s (\\citeyear{ab99}) diagnostics, they found bar signatures in $14$ of $17$ galaxies with a B/PS bulge ($\\approx80\\%$) and in none of the galaxies with a more spheroidal bulge. In spite of the aforementioned studies, our knowledge of the stellar kinematics of B/PS bulges and their hosts remains limited. This probably arises because the gas is a more practical tracer of the barred potential, being both easier to observe and reacting more strongly than the stars to asymmetries (shocks). Studying the stars directly is nevertheless essential as B/PS bulges are stellar structures and, when comparing to N-body simulations, stellar kinematics largely bypasses issues such as the selection of an optimal gaseous tracer (H$\\alpha$, CO, \\ion{H}{1}, etc), star formation, and (for gas-poor objects) the response of the orbital structure to a massive gas disk \\citep[e.g.][]{bhsf98}. Existing stellar kinematic studies of B/PS bulges are limited to a few objects and have largely focused on the issue of cylindrical rotation, using slits offset from the major-axis or perpendicular to it \\citep*[e.g.][]{bc77,ki82,r86,j87,s93a,swc93}. As noted above, however, this is consistent with both axisymmetric and barred configurations, and so does not discriminate between the two dominant formation scenarios. We advocate here instead the new kinematic bar diagnostics of \\citet{ba04}, which directly probe the shape of the potential \\citep[but see also][]{bg94,km95}. The importance of B/PS bulges goes beyond their sheer number and interesting morphology, as they may hold vital clues for our understanding of bulge formation. There is growing evidence against significant merger growth in many bulges (e.g.\\ \\citealt*{k93,apb95,mch03,bgdp03}) and a corresponding emphasis on secular evolution processes, largely bar-driven \\citep[e.g.][]{wfmmb95,es02}. Most theoretical models involve the growth of a central mass through bar inflow and/or possibly recurring bar destruction \\citep*[e.g.][]{pn90,fb93,fb95,nsh96}, although the efficiency of bar dissolution mechanisms remains uncertain (\\citealt{ss03} and references therein). Nevertheless, because of their probable relationship to bars, B/PS bulges are likely to play a central role in those scenarios, and a proper understanding of their structure and dynamics is essential to constrain them. Our primary goal in this work is thus to study the stellar kinematics of a statistically significant number of galaxies with a B/PS bulge, simultaneously and independently verifying their relationship to bars and probing for embedded structures. We study \\citeauthor{bf99}'s (\\citeyear{bf99}) sample of $23$ edge-on spirals, $17$ with a B/PS bulge and $6$ with other bulge types, as well as an additional $7$ B/PS bulges which showed little (or confined) emission and were only briefly discussed by the authors. The latter offer the best comparison with N-body models, being largely gas and dust free. In \\S~\\ref{sec:diagnostics}, the stellar kinematic bar diagnostics of \\citet{ba99} and \\citet{ba04} are summarized. In \\S~\\ref{sec:obs}, we describe the sample, observations, as well as the reduction and analysis of the data. The stellar kinematics of the $30$ galaxies is presented briefly in \\S~\\ref{sec:results}, while \\S~\\ref{sec:discussion} discusses in greater detail the structure and dynamics of B/PS bulges and how they fit into bar-driven secular evolution scenarios. We summarize our results and conclude briefly in \\S~\\ref{sec:conclusion}. ", "conclusions": "} The (projected) morphology of B/PS bulges departs strongly from that of classical, spheroidal bulges \\citep[e.g.][]{s93b}, implying an even more distorted 3D structure, but their high incidence ($\\gtrsim45\\%$ of all bulges; see \\citealt{ldp00a}) makes them more than simple curiosities and a widespread mechanism must be able to account for their structure and dynamics. By studying the stellar kinematics of the $30$ edge-on spiral galaxies of \\citet{bf99}, we have attempted here to discriminate between the two leading formations scenarios: accretion \\citep[e.g.][]{bp85,r88} and bar-buckling \\citep[e.g.][]{cs81,cdfp90}. The latter is particularly attractive in view of currently popular secular evolution scenarios, which argue that bars may drive bulges along the Hubble sequence (e.g.\\ \\citealt{fb95}; \\citealt{nsh96} and references therein). Our results are summarized in Table~\\ref{tab:class}. $22$ of $24$ galaxies with a B/PS bulge show clear kinematic bar signatures ($\\approx90\\%$), while only $2$ of the $6$ control sample galaxies do and are probably transition objects. We thus confirm the results of \\citet{bf99} for galaxies with extended ionized-gas and, for the first time, extend those results to gas-poor B/PS bulges (S0s). The kinematic evidence in favor of a close relationship between thickened bars and B/PS bulges is thus now overwhelming across all morphological types (S0--Sbc), although we still have not probed the potentials out of the equatorial plane. Zamojski et al.\\ (2004, in preparation) shall discuss the stellar kinematics of a few objects at large galactic heights. Most objects show kinematic profiles in agreement with those predicted from N-body simulations of barred disks \\citep{ba04}, including the expected relationship between bulge morphology and bar viewing angle. In particular, most galaxies with a B/PS bulge have i) a double-hump rotation curve with a dip or plateau at moderate radii, ii) a rather flat central velocity dispersion profile with a secondary peak or plateau and (in about $40\\%$ of cases) a local central minimum, and iii) an $h_3$ profile correlated with $V$ over the expected bar length, with up to three slope reversals. As expected if caused by a bar, the strengths and extents of those features are correlated. Our $h_4$ profiles are generally too noisy to be of much use. A number of differences with the simulations also exist, most of which appear related to the past or current presence of gas. In particular, the kinematic profiles of most galaxies show an anti-correlation of $V$ and $h_3$ in the very center, over a region corresponding to the rapidly rising part of the rotation curve (or at most the first plateau) and encompassing the central light peak and central $\\sigma$ minimum (if present). Being cospatial with observed ionized-gas disks in gas-rich galaxies \\citep{bf99}, this kinematic behavior is most likely due to cold quasi-axisymmetric central stellar disks, which we argue have decoupled due to bar-driven gas inflow, and may have been enhanced by subsequent star formation. B/PS bulges are thus entirely consistent with the predictions of most bar-driven secular evolution scenarios \\citep[e.g.][]{hs94,fb95}, whereby bars can simultaneously account for the creation of a large vertically extended component out of disk material and for the accumulation of mass in their centers (possibly associated with the so-called bulges of face-on systems), bypassing the need for merging or accretion of external material. Lastly, we reiterate that $h_3$ is a very useful tracer of the structure and dynamics of galaxies. As suggested by \\citet{ba04}, the correlation or anti-correlation of $h_3$ and $V$ appears to be a reliable diagnostics of the intrinsic shape, at least for highly inclined disks." }, "0403/astro-ph0403281_arXiv.txt": { "abstract": "We confirm the orbital period of WX~Cen $\\equiv$ WR~48c determined by \\citet{dia95} -- DS95 -- and refined its value to $P_{orb} = 0.416~961~5 (\\pm 22)$ d. The light curve of this object has a peak to peak variation of approximately 0.32 magnitudes. It is non-sinusoidal in the sense that it has a V-shaped narrow minimum, similar to the ones seen in V~Sge, V617~Sgr and in Compact Binary Supersoft Sources -- CBSS. Most of the emission lines in the optical spectrum are due to Balmer, He II, C IV, N V, O V and O VI. An analysis of the He II Pickering series decrement shows that the system has significant amount of hydrogen. The emission lines of He II 4686{\\AA} became weaker between the 1991 and 2000/2002 observations, indicating distinct levels of activity. The spectra of WX~Cen show variable absorption features in the Balmer lines with $V = -2900$ km~s$^{-1}$ and in emission with $V = \\pm 3500$ km~s$^{-1}$. These highly variable events remind the satellites in emission of CBSS. We estimate the color excess as $E(B-V)=0.63$ on the basis of the observed diffuse interstellar band at 5780{\\AA}. Given the distance-color excess relation in the direction of WX Cen, this implies a distance of $2.8 \\pm 0.3$ kpc. Interstellar absorption of the Na I D lines show components at $-4.1$ km~s$^{-1}$, which corresponds to the velocity of the Coalsack, and three other components a $-23.9$, $-32.0$ and $-39.0$ km~s$^{-1}$. These components are also seen with similar strengths in field stars that have distances between 1.8 and 2.7 kpc. The intrinsic color of WX Cen is $(B-V)_0=-0.2$ and the absolute magnitude, $M_V = -0.5$. Extended red wings in the strong emission lines are seen. A possible explanation is that the system has a spill-over stream similar to what is seen in V617~Sgr. We predict that when observed in opposite phase, blue wings would be observed. A puzzling feature that remains to be explained is the highly variable red wing ($V \\sim 700$ km~s$^{-1}$) of the O VI emission lines as well as of the red wings of the H and He lines. The velocity of the satellite-like feature is consistent with the idea that the central star is a white dwarf with a mass of $M \\sim 0.9 M_{\\sun}$. With the high accretion rate under consideration, the star may become a SN~Ia in a time-scale of $5 \\times 10^{6}$ years. ", "introduction": "WX~Cen was initially identified by \\citet{egg} as a possible optical counterpart of the hard X-ray transient source Cen XR-2, although this identification was latter discarded. Because of its spectral characteristics, the object was then classified as a Wolf-Rayet star of type WN 7 \\citep{huc81}, while \\citet{vogt} classified it as a nova-like variable. \\citet{dia95} -- DS95 -- showed that it is a binary system with an orbital period of 10.0 hr. The photometric orbital variation, determined by these authors, based on their spectroscopic observations, is approximately sinusoidal with an amplitude of $\\sim 0.3$ mag. In spite of the period being relatively long (in the context of Cataclysmic Variables) it was not possible to identify any spectral signature of the secondary star. A distance of 1400 pc was determined from the Na I D line equivalent width, and an $E(B-V) = 0.4 \\pm 0.1$ was suggested from the average of three distinct determinations (4430{\\AA} DIB and Na I D equivalent widths and He II 4686, 10124 {\\AA} lines ratio). Doppler tomography of WX~Cen produced by DS95 suggests that the gas of the secondary has normal chemical abundance, while the wind of the primary star is probably over-abundant in helium (He$^{++}$/H = 0.56). DS95 also showed that, if the primary component of WX~Cen is a white dwarf, then the secondary should have a mass bellow $0.35 M_{\\sun}$ and, therefore, be evolved. If, however, one assumes the hypothesis that the secondary is a main sequence star, it would have $M_2 = 1.16 M_{\\sun}$, and the primary, a mass of $M_1 > 3.5 M_{\\sun} \\pm 0.5 M_{\\sun}$. \\citet{steiner98} included WX~Cen in a group of 4 galactic binaries defined as the V~Sge stars. The other three objects are V~Sge \\citep{her,dia99}, V617~Sgr \\citep{steiner99,cie} and DI~Cru \\citep{vee02c}. They are characterized by the presence of strong emission lines of O VI and N V. Besides, He II is at least two times more intense than H$\\beta$. The V~Sge stars are similar to the Compact Binary Supersoft Sources (CBSS), seen in the Magellanic Clouds, but not that frequent in the Galaxy. CBSSs are interpreted as suffering hydrostatic hydrogen nuclear burning on the surface of a white dwarf. This burning is due to the high mass-transfer rate, a consequence of the fact that the systems have inverted mass ratios (see \\citet{kah} for a revision and references). \\citet{patt} also considered the hypothesis of Galactic CBSS nature for the stars V Sge, T Pyx and WX Cen, established on their extremely blue colors, high luminosities, orbital light curves and highly excited emission line spectra. In an observational program to search for galactic WR stars, \\citet{shara} rediscovered WX~Cen as a new WN3, when this star received the WR~48c designation. In our search for V~Sge type stars, we selected WR~48c as a candidate, on the basis of spectroscopic criteria, for a detailed observational study. We, then, realized that WR~48c was the already known star WX~Cen. \\begin{figure} \\vspace*{10pt} \\centerline{\\includegraphics[width=84mm]{fg1.eps}} \\caption{Average \\textit{V} light curve of the data obtained in 2000, folded with the orbital period and epoch from the photometric ephemeris. \\label{lca}} \\end{figure} \\begin{figure} \\vspace*{50pt} \\centerline{\\includegraphics[width=84mm]{fg2.eps}} \\caption{The light curve of WX~Cen for the nights 2000 April 24, 25 and 26. The upper light curves are of the comparison star. An increase of 0.08 mag in the mean magnitude is seen in consecutive nights. \\label{lcb}} \\end{figure} ", "conclusions": "The main conclusions of this paper are: \\begin{enumerate} \\item We confirmed the orbital period of WX~Cen determined by DS95 and refined its value to $P_{orb} = 0.416~961~5 (\\pm 22)$ d, based on spectroscopic observations. \\item The light curve has total amplitude of approximately 0.32 magnitudes and is non-sinusoidal in the sense of having a narrow, V-shaped, minimum. The object presents flickering with time-scales of tens of minutes. Night-to-night variations of about 0.08 mag are also observed. \\item We identified most of the emission lines as due to Balmer, He II, C IV, N V, O III, O V, O VI. \\item The object shows absorption satellites in the Balmer lines with $V = -2900$ km~s$^{-1}$ and in emission with $V = \\pm 3500$ km~s$^{-1}$. These highly variable events remind the satellites in emission in CBSS. \\item The analysis of the emission lines show that the He II 4686{\\AA} line became weaker between 1991 and 2000/2002 observations. At the same time the emission lines were narrower, suggesting that in 2000/03 the system was in a less active state than in 1991. \\item The object presents red-shifted emission in high ionization species. Extended red wing emission is also seen in the strongest lines. The O VI lines show strong variability with velocity of about 700 km~s$^{-1}$. \\item An analysis of the He II Pickering series decrement shows that the system has significant amount of hydrogen, with an abundance, by number, between 2.1 and 4.5. \\item We estimate the color excess as $E(B-V)=0.63$ on the basis of the observed DIB at 5780{\\AA}. Given the distance-color excess relation in the direction of WX Cen, this implies a distance of $2.8 \\pm 0.3$ kpc. Interstellar absorption of the Na I D lines show components at $-4.1$ km~s$^{-1}$, which corresponds to the velocity of the Coalsack, and three other components a $-23.9$, $-32.0$ and $-39.0$ km~s$^{-1}$. These components are also seen with similar strengths in field stars that have distances between 1.8 and 2.7 kpc. The intrinsic $B-V$ color index of WX Cen is $(B-V)_0=-0.2$ and the absolute magnitude, $M_V = -0.5$. \\item A possible explanation for the extended wings in the strong emission lines is that the system has a spill-over stream similar to what is seen in V617~Sgr. We predict that when observed in opposite phase, blue wings would be observed. \\item The velocity of the satellites is consistent with the idea that the central star is a white dwarf with a mass of $M \\sim 0.9 M_{\\sun}$. With the high accretion rate, it may become a SN~Ia in a time-scale of $5 \\times 10^{6}$ years. \\end{enumerate}" }, "0403/astro-ph0403248_arXiv.txt": { "abstract": "We have completed an intensive monitoring program of two fields on either side of the center of M31, and report here on the results concerning microlensing of stars in M31. These results stem from a three-year study (the VATT/Columbia survey) of microlensing and variability of M31 stars, emphasizing microlensing events of 3 day to 2 month timescales and likely due to masses in M31. These observations were conducted intensively from 1997-1999, with baselines 1995-present, at the Vatican Advanced Technology Telescope and the 1.3-meter telescope at MDM Observatory, with additional data from the Isaac Newton Telescope, including about 200 epochs total. The two fields monitored cover 560 square arcminutes total, positioned along the minor axis on either side of M31. Candidate microlensing events are subject to a number of tests discussed here with the purpose of distinguishing microlensing from variable star activity. A total of four probable microlensing events, when compared to carefully computed event rate and efficiency models, indicate a marginally significant microlensing activity above that expected for the stars alone in M31 (and the Galaxy) acting as lenses. A maximum likelihood analysis of the distribution of events in timescale and across the face of M31 indicate a microlensing dark matter halo fraction consistent with that seen in our Galaxy towards the Large Magellanic Cloud (Alcock et al.~2000a). Specifically, for a nearly singular isothermal sphere model, we find a microlensing halo mass fraction $f_b=0.29^{+0.30}_{-0.13}$ of the total dark matter, and a poorly constrained lensing component mass (0.02 to $1.5 M_\\odot$, 1 $\\sigma$ limits). This study serves as the prototype for a larger study approaching completion; between the two there is significant evidence for an asymmetry in the distribution of microlensing events across the face of M31, and therefore a large population of halo microlensing dark matter objects. ", "introduction": "\\bigskip The nature of the dark matter in the halo of disk galaxies remains a mystery. While this component of galaxies contributes the majority of their mass, most other data about the dark matter are indirect at best. Some clue as to characteristics of part of the dark matter may be implied by the microlensing detection of masses towards the Large Magellanic Cloud. The MACHO survey revealed a frequency of microlensing that was unanticipated in the context of the visible stellar population alone, but still falls short of accounting for the entire dark matter halo (Alcock et al.~2000a). In particular, the survey indicates a most probable dark matter halo mass fraction of 20\\% (limits of 5\\% to 50\\%, at the 95\\% confidence level), for an indicative spatial distribution model (singular, isotropic, isothermal sphere). Constraints from the EROS magellanic cloud survey (Afonso et al.~2003) yields a constraint consistent with 20\\% microlensing halo fraction, but also consistent with no microlensing halo. There are few independent indications that bear on the validity of the presence of a microlensing halo component. Fluctuations in the brightnesses of images of lensed QSO MG~0414+0534 indicate a significant but sub-dominant halo fraction of microlensing masses (Schechter \\& Wambsganss 2002). A decade ago we proposed that M31 offers a favorable alternative venue for probing the halo dark matter problem in spiral galaxies, by applying the microlensing techniques to stars in M31 itself, for lenses primarily in M31 but also the Galaxy. In particular, such a signal could be easily distinguished in terms of an asymmetry in microlensing rate across the face of M31, and could be monitored effectively using image subtraction to suppress the severe crowding of M31's stars (Crotts 1992). This required developing techniques for the application of image subtraction to a time series of images (Tomaney \\& Crotts 1996), which led to the first candidate microlensing events in M31 (Crotts \\& Tomaney 1996). At least several more years of such observations have been required to both amass sufficient lensing events for a statistically meaningful sample and to cull out variable stars by extending the baseline. The current work presents the results from this more extended survey and offers our interpretation of the microlensing observations which have resulted. While our survey has found several thousand variable stars, to be reported elsewhere, we have also isolated a sample of events that are more consistent with microlensing events, and imply that a significant fraction of the dark matter halo in M31 is due to objects of stellar mass. M31 microlensing surveys have also been conducted by the AGAPE groups (Baillon et al.~1993, Calchi Novati et al.~2002), and the survey reported here has been extended by MEGA (Crotts et al.~2001, deJong et al.~2003), the first results of which we mention in conjunction with these in reaching our conclusion. ", "conclusions": "How do our results compare to those from Galactic Halo microlensing searches? The most detailed results come from the MACHO survey, reporting (Alcock et al.~2000a) that, during 5.7 years of observations, the detection of 13-17 events towards the LMC when only $\\sim$2-4 events would have been expected due to known, intervening stellar populations. This corresponds to a microlensing optical depth, $\\tau_{LMC} = 1.2_{-0.3}^{+0.4} \\times 10^{-7}$ for $2 < \\hat{t} < 400$ days. The events have timescales of $<\\hat{t}>$ = 34-230 days and the most probable mass for the events is $$ = 0.15-0.9 \\msolar. The maximum likelihood halo fraction is $f_{b} $ = 20\\% (8\\% - 50\\%, 95\\% confidence interval) implying a total mass in MACHOs of M$_{\\rm MACHO}$ = $9_{-3}^{+4} \\times 10^{10}$ \\msolar~ ($r <$ 50 kpc). The EROS survey reports a result consistent with the central MACHO value, but also consistent with no Galactic microlensing halo (Afonso et al.~2003). In addition to searching for events with well-sampled, long-duration microlensing lightcurves, the MACHO \\& EROS groups conducted an analysis (``spike analysis'') in which they searched for very short timescale brightenings in order to place limits on low mass MACHOs. Their conclusion is that Milky Way halo dark matter cannot be comprised of objects in the mass range $2.4 \\times 10^{-7}$ \\msolar $< m <$ $5.2 \\times 10^{-4} M_\\odot$ (Alcock et al.~1996). The M31 halo microlensing fraction $f_b$ we find is consistent with that seen by MACHO towards the LMC. Our central value is higher (by $0.7\\sigma$ using only our error, or $\\approx 0.4\\sigma$ combining both surveys' errors). We cannot argue for any inconsistency between the two surveys' results. The positive, marginally significant halo signal we report is due in large part to the asymmetric distribution of events across the face of M31, slightly favoring the far side as would be expected from a microlensing dark matter halo. Our result is approximately as consistent with no halo as that reported by MACHO, however. On the basis of further M31 microlensing evidence, however, we tend to accept the positive halo indication. The VATT/Columbia survey serves as the pilot study for a larger survey, MEGA (``Microlensing Exploration of the Galaxy and Andromeda:'' Crotts et al.~2001), and the first results from this effort (de Jong et al.~2003) also shows a marginally significant farside surplus asymmetry. While this other work does not estimate $f_b$, such an asymmetry is a nearly unique marker of a dark matter halo of microlensing objects. Later data from this survey will encompass almost an order of magnitude more observations, so indicates that the result will become more clear in the near future. Perhaps the most surprising conclusion drawn from the MACHO data is that the lenses lie in a mass range occupied by stellar objects and are well above the hydrogen burning limit ($\\sim$0.065 \\msolar: Chabrier \\& Baraffe [2000]) which defines the brown dwarf/stellar boundary. Creating plausible models which can explain these results without violating other astrophysical constraints has proven to be quite a challenge for theorists. The only candidates in this mass range to be excluded by direct observation are low-mass stars (Boeshaar, Tyson \\& Bernstein 1994; Graff \\& Freese 1996) and the most viable baryonic candidates are a population of old white dwarfs (WD) which have been previously overlooked because their colors are bluer than expected due to an increase in atmospheric H$_{2}$ opacity (Hansen 1999). Various surveys (Ibata et al.~1999, 2000; Oppenheimer et al.~2001a, b) have now discovered a significant number of such high-proper motion, faint WDs whose kinematics appear to be consistent with those of Galactic halo objects. Alternatively, there are those who would argue that these events are not associated with a dark matter halo but are, rather, explained as lensing by an intervening population of stars along the line-of-sight to the Magellanic Clouds (see Zaritsky \\& Lin 1997, Zhao 1998) or as self-lensing by stars within the LMC/SMC (Sahu 1994). As an alternative, we point out that primordial black holes with masses of $\\sim$1 \\msolar\\, can be formed from density perturbations created during the quark-hadron phase transition in the early universe (Crawford \\& Schramm 1982, Jedamzik 1997). However, although initially baryonic these black holes are classified as non-baryonic dark matter candidates because they form before the epoch of BBN and, therefore, are not subject to its constraints. Comparing the current survey to MACHO, their constraints on $m$ is somewhat more restrictive. They argue for masses above the 0.07$M_\\odot$ hydrogen core-burning threshold, at a level of certainty of about 90\\%. While in M31 we cannot rule out the possibility that this represents a population of red main-sequence dwarf stars, in our own Galaxy this is ruled out. While our current survey adds little new information as to the nature of these objects, it tends to confirm that the population is seen to exist now in two galaxies, and is likely to be a universal phenomenon. With galaxy halo dark matter accounting for nearly the same fraction of universal closure density as baryons (Rubin 1993), with the WMAP value now set at $\\Omega_B = 0.046$ (Spergel et al.~2002). We are speaking of a contribution of about 1\\%, well within the baryon budget. An M31 microlensing result might be contaminated by foreground lenses, as has been suggested for the LMC. Since the timescales of Galactic and M31 halo events are similar, given the same lens mass, an M31 survey is susceptible to foreground Galactic halo lensing. Extrapolating previous results (Zhao 1998), an optical depth of $10^{-7}$ with a sheet of matter at 10 kpc from the source (or observer), implies a surface density $\\sim 15 M_\\odot/pc^2$. Our signal implying a microlensing halo corresponds to an optical depth roughly an order of magnitude larger than this. In the Milky way, just judiciously covering the MDM farside field with such a sheet would only require $3\\times 10^5M_\\odot$. In M31, this requires $\\sim 10^9 M_\\odot$. The latter would imply an unusually massive tidal stream, whereas such a possibility in our Galaxy is perhaps reasonable if one ignores the double coincidence of its appearance both in front of the MACHO and current M31 survey fields. For either galaxy, a spherical shell would need about $\\sim 10^{11}M_\\odot$ at 10 kpc to make the optical depth, with mass at other locations scaling as the radius: $M_{shell} \\approx 10^{11} M_\\odot ( D / 10~kpc )$. This approaches a simple re-creation of the original halo lensing mass problem. From this survey we have insufficient data to state whether these lenses arise in a thick disk or a true halo. To do so will require more events, scattered over a larger portion of the face of M31. Fortunately, such a survey is practible (Baltz et al.~2003) and is currently underway (Crotts et al.~2001). While the results presented here are interesting in their implications for the universality of disk galaxy halo dark matter, we would prefer to have better sampled lightcurves, to cover more of the face of M31 in order to get better leverage on the spatial variation of the lensing population, and to better catalog populations of variable stars in the same field which might masquerade as microlensing events. We look forward to additional progress in using M31 to generalize and extend the insight which has been won from microelensing searches in nearby galaxies." }, "0403/astro-ph0403412_arXiv.txt": { "abstract": "We present $\\sim$2$''$-4$''$ aperture synthesis observations of the circumstellar disk surrounding the nearby young star TW Hya in the CO J=2--1 and J=3--2 lines and associated dust continuum obtained with the partially completed Submillimeter Array. The extent and peak flux of the 230 and 345 GHz dust emission follow closely the predictions of the irradiated accretion disk model of \\citet{calvet_d02}. The resolved molecular line emission extends to a radius of at least 200 AU, the full extent of the disk visible in scattered light, and shows a clear pattern of Keplerian rotation. Comparison of the images with 2D Monte Carlo models constrains the disk inclination angle to $7^{\\circ}\\pm 1^{\\circ}$. The CO emission is optically thick in both lines, and the kinetic temperature in the line formation region is $\\sim$20~K. Substantial CO depletion, by an order of magnitude or more from canonical dark cloud values, is required to explain the characteristics of the line emission. ", "introduction": "TW Hya is the closest known classical T Tauri star, and exhibits high X-ray flux, large lithium abundance, and evidence for an actively accreting disk with accretion rate estimated to be from 10$^{-9}$ to 10$^{-8}$ M$_{\\odot}$ yr$^{-1}$ based on the large H${\\alpha}$ equivalent width and the X-ray emission (\\citealp{kastner_h02} and refs therein). Despite its apparently advanced age, estimated to be 5 to 20 Myr, TW Hya is surrounded by a disk of mass $\\sim 5 \\times 10^{-3}$ M$_{\\odot}$ inferred from dust emission (\\citealp{wilner_h00}). The disk is viewed nearly face-on, and is visible in scattered light to a radius of at least $3\\farcs5$, or 200 AU at a distance of 56 pc (\\citealp{krist_s00, trilling_k01, weinberger_b02, calvet_d02}). Several molecular species (CO, HCN, CN, HCO$^+$, DCO$^+$) have been detected in the TW Hya disk at millimeter and submillimeter wavelengths using single dish telescopes (\\citealp{kastner_z97, vanzadelhoff_v01, thi01, vandishoeck_t03}). The proximity, isolation, and rich chemistry of the TW Hya disk make this system an excellent target for the study of the physical and chemical structure of protoplanetary environments at high spatial resolution using interferometry. Models may be directly compared to aperture synthesis images of molecular lines to verify the inferences from low resolution observations, and to further probe the kinematics, physical conditions, and chemistry. A first step in this direction was the imaging analysis of HCO$^+$ J=1--0 emission observed with the Australia Telescope Compact Array by ~\\citet{wilner_b03}, though the spatial and spectral resolution were insufficient to examine the kinematics in any detail. The prospects are excellent for interferometric observations of protoplanetary disks at submillimeter wavelengths (\\citealp{blake02}). The Submillimeter Array (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.} , now nearing completion on Mauna Kea, is an ideal instrument for making these observations (\\citealp{ho_m04}). In this {\\em Letter}, we present early SMA observations of the TW Hya disk in the CO J=3--2 and J=2--1 lines and dust continuum, with a maximum resolution of $2''$. The resulting images unambiguously reveal the rotation of the disk around the star and constrain the disk inclination and size. ", "conclusions": "" }, "0403/astro-ph0403138_arXiv.txt": { "abstract": "Gamma-ray spectra from cosmic-ray proton and electron interactions with dense gas clouds have been calculated using a Monte Carlo event simulation code, GEANT4. Such clouds are postulated as a possible form of baryonic dark matter in the Universe. The simulation fully tracks the cascade and transport processes which are important in a dense medium, and the resulting gamma-ray spectra are computed as a function of cloud column-density. These calculations are used for predicting the Galactic diffuse gamma-ray spectrum which may be contributed by baryonic dark matter; the results are compared with data from the EGRET instrument, and used to constrain the fraction of Galactic dark matter which may be in the form of dense gas clouds. In agreement with previous authors, we find useful constraints on the fraction of Galactic dark matter which may be in the form of low column-density clouds ($\\Sigma\\la 10\\,{\\rm g\\,cm^{-2}}$). However, this fraction rises steeply in the region $\\Sigma\\sim10^2\\,{\\rm g\\,cm^{-2}}$, and for $\\Sigma\\ga200\\,{\\rm g\\,cm^{-2}}$ we find that baryonic dark matter models are virtually unconstrained by the existing gamma-ray data. ", "introduction": "The nature of dark matter remains one of the outstanding questions of modern astrophysics. The success of the cold dark matter cosmological model (albeit with ``dark energy'' now required: $\\Lambda$CDM) argues strongly for a major component of the dark matter being in the form of an elementary particle. However, the inventory of baryons which we can observe locally falls far short of the total inferred from observations of the cosmic microwave background fluctuations \\citep{Fuk04}, leaving open the possibility that there may be a significant baryonic component of dark matter. Furthermore, although $\\Lambda$CDM is very successful in describing the growth of structure in the universe on large scales, we still lack a direct detection of any of the candidate dark matter particles. Lacking this decisive piece of observational evidence, some authors have proposed models which include a large component of baryonic dark matter. In particular there have been many papers dealing with the possibility that cold, self-gravitating molecular clouds constitute a major component of the dark matter \\citep{Pfe94,DeP95,Hen95,Ger96, Com97, Wal99, Sci00a, Sci00b}. A variety of different forms, including isolated, clustered, and fractal, have been considered for the clouds, but all proposals involve dense gas of high column-density, in contrast to the diffuse gas in the interstellar medium which is easily detected in emission and/or absorption. One of the fundamental predictions of a model featuring dense gas clouds is the gamma-ray emission resulting from cosmic-ray interactions within the clouds \\citep{DeP95, Kal99, War99, Sci00a}. Because of the potentially large total mass of gas involved, this process may yield a diffuse flux in the Galactic plane comparable to the flux from known sources for photon energies around 1 GeV \\citep{Sci00a}. Considering the high quality data on diffuse emission acquired by the EGRET detector aboard the Compton Gamma Ray Observatory \\citep{Hun97}, it is worth considering this source of gamma-ray emission in detail as it is possible to use these data to constrain the dark matter models (see \\citet{Sal96}; \\citet{Gil94}). Most previous investigations of this problem have neglected the self-shielding and cascade phenomena which can be important at high column densities \\citep{Kal99, Sci00a}, and have employed emissivities appropriate to the low-density limit. These effects alter the emergent gamma-ray spectrum, and we note that this could be relevant to the observed excess Galactic flux above 1 GeV \\citep{Hun97}. We have noted elsewhere \\citep{Wal03} that massive ($M\\ga10^6\\;{\\rm M_\\odot}$) aggregates of dense gas clouds could potentially account for many of the unidentified discrete sources detected by EGRET \\citep{Har99}. Here we present detailed calculations of the gamma-ray spectra arising from cosmic-ray interactions with dense gas clouds. We have used a Monte Carlo simulation code, GEANT4, developed for simulating interaction events in detectors used in high-energy particle physics. Not surprisingly, we find that the predicted spectra differ substantially between high and low column-density clouds, and we discuss the interpretation of our results in the context of the observed Galactic gamma-ray emission. Our calculations are undertaken for cold, dense molecular gas in clouds of radius $R\\sim 10^{13}$cm, similar to those proposed by \\citet{Wal98} to explain the extreme scattering events \\citep{Fie94} during which compact extragalactic radio sources are magnified and demagnified as a plasma ``lens'' moves across the line of sight (see \\citet{McK01} for a criticism of this model). However, the results of our calculations depend primarily on the column-density of the individual clouds, $\\Sigma$, under consideration, and their fractional contribution to the Galaxy's dark matter halo, and our results can be taken as representative of other models which are characterised by similar values of these quantities. ", "conclusions": "\\subsection{Comparison with data} Several physical processes contribute to the observed diffuse gamma-ray intensity. In order of decreasing fractional contribution these are thought to be pion production and bremsstrahlung from cosmic-ray interactions with {\\it diffuse\\/} Galactic gas, inverse Compton emission from cosmic-ray electrons interacting with ambient photons, and an isotropic background, which is presumably extragalactic and due to many faint, discrete sources. The sum of these contributions offers a good model for the diffuse emission which is actually observed in the 100~MeV--1~GeV band \\citep{Blo89, Hun97}; however, the $E>\\,1$~GeV emission is poorly modeled \\citep{Hun97} with the prediction \\citep{Ber93} amounting to only $\\sim2/3$ of the observed intensity. Some authors \\citep{Gil94, Sal96} have used the low-energy ($E<\\,1$~GeV) data to argue that the agreement between model and data allows no room for any significant unmodeled emission and that these data therefore place tight constraints on any baryonic contribution to the dark matter halo of the Galaxy. This line of argument is clearly suspect because of the failure of the same model to account for the high-energy ($E>\\,1$~GeV) data. However, even if we ignore the high-energy data the calculations presented in \\S2.3 demonstrate that the gamma-ray constraints on high column density gas clouds in the dark halo are much weaker than those on low column density gas because the emissivity of the latter is much greater; consequently, we revisit the published constraints in the following section (\\S3.2), returning to the question of the high-energy data in \\S3.3. \\subsection{Published constraints applied to high column-density gas} \\citet{Gil94} argued that the known (diffuse) gas accounts for essentially all of the gamma-ray intensity in the $E>100$~MeV band observed by the COS B satellite \\citep{Blo89}, and he suggested that any contribution from baryonic clouds should amount to no more than $10^{-5}\\,{\\rm ph\\,cm^{-2}s^{-1}sr^{-1}}$ ($E>100$~MeV) at high Galactic latitudes. Toward the Galactic poles we found (\\S3) $Q\\simeq2.5\\times10^{-3}{\\rm g\\,cm^{-2}}$, implying that the average emissivity of the material in the dark halo should be ${\\cal E} < 5.0\\times10^{-2}\\,{\\rm ph\\,s^{-1}g^{-1}}$ for $E>100\\;{\\rm MeV}$. From Table 1 (or figures 6 and 7) we see that this requirement is not met for the models which are in the thin material limit but that any model with $\\Sigma\\ga100\\,{\\rm g\\,cm^{-2}}$ is acceptable. In other words the constraint imposed by \\citet{Gil94} permits the Galaxy's dark halo to be entirely baryonic, provided the individual components have a column-density $\\Sigma\\ga100\\,{\\rm g\\,cm^{-2}}$. \\citet{Sal96} pointed out that obervations at low Galactic latitudes can provide tighter constraints on any baryonic dark halo, because of the higher intensity expected when looking through the cosmic-ray disk edge-on. This point is manifest in the large values of $Q(b=0)$, which we computed in \\S2.3 (see also figure 8). \\citet{Sal96} argued that a suitable constraint on any baryonic component of the dark halo is that it should not contribute more than the uniform background intensity component observed in any given field. The strongest constraint then comes from observations of the Ophiuchus region \\citep{Hun94}, for which the requirement imposed by \\citet{Sal96} is an intensity contribution from the dark matter halo of $I_D<2.4\\times10^{-6}{\\,\\rm ph\\,cm^{-2}s^{-1}sr^{-1}}$ in the band 300\\ MeV $\\le E \\le$ 500\\$ MeV. For this line of sight our calculation yields $Q(0, b=15^\\circ)=1.66\\times10^{-2}{\\rm g\\,cm^{-2}}$, and in the thin material limit, for which we find ${\\cal E}\\simeq2.4\\times10^{-2}\\,{\\rm ph\\,s^{-1}g^{-1}}$ (300\\ MeV $\\le E \\le$ 500\\$ MeV; see Table 1), this corresponds to a predicted intensity of $I_D\\simeq3.2\\times10^{-5}{\\,\\rm ph\\,cm^{-2}s^{-1}sr^{-1}}$ in the 300--500~MeV band, implying that at most 8\\% of the dark halo can be resident in low column density clouds.\\footnote{This result is a factor of 2 larger than the corresponding model of \\citet{Sal96} (with $L=3$~kpc and a spherical halo), a difference which is accounted for by their choice of a smaller core radius for the dark halo. \\citet{Sal96} chose a core radius of 3.5~kpc, which yields $Q(0, b=15^\\circ)=3.0\\times10^{-2}{\\rm g\\,cm^{-2}}$, whereas we have employed a core radius of 6.2~kpc (see \\S2.3).} However, our calculations extend to clouds of higher column-densities, and we find that for $\\Sigma\\ga500\\,{\\rm g\\,cm^{-2}}$ the limit relaxes to the point where all of the dark halo is permitted to be in the form of high column density clouds. \\subsection{Revision of constraints based on the observed spectrum} \\subsubsection{Galactic diffuse gamma-ray model} The constraints discussed in the previous subsection make reference only to the low-energy ($E<\\,1$~GeV) gamma-ray data. As noted earlier in this section, the observed intensity of the inner Galactic disk at high photon energies ($>\\,1$~GeV) is substantially greater than expected \\citep{Hun97, Ber93}. Much effort has been expended on explaining this discrepancy, with most of the attention given to models in which the Galactic cosmic-ray electron and/or proton spectra differ from their locally measured values \\citep{Bus01, Aha00, Ber00, Str00, Mor97}. However, none of these models offers a satisfactory explanation of the observed mean gamma-ray spectrum of the Galactic disk, and consequently a successful match to the low energy data should not be taken to mean that the emission model is correct. In turn this suggests that the constraints formulated by \\citet{Gil94} and \\citet{Sal96}, on the basis of the low-energy data, may be too restrictive. The question then arises as to what constraints the gamma-ray data do in fact place on unmodeled emission, given the current state of understanding of the observed emission. The Galactic diffuse emission model used in \\citet{Hun97} is based on \\citet{Ber93}, and contains the following contributions: \\begin{equation} I = I_{\\rm HI}+I_{\\rm HII}+I_{\\rm H_2} + I_{\\rm IC}+I_{\\rm EG}, \\end{equation} where $I_{\\rm HI}$ is the gamma-ray intensity contributed by cosmic-ray interactions with diffuse atomic hydrogen, and similarly for the ionised and molecular components of the interstellar medium (HII and ${\\rm H_2}$, respectively). The emissivity for these components \\citep{Ber93} is, of course, computed in the thin material limit, and each component therefore has the same spectral shape, differing only in normalisation. Here $I_{\\rm IC}$ is the gamma-ray flux by inverse Compton emission (cosmic-ray electrons up scattering low-energy photons), and $I_{EG}$ is the isotropic extragalactic background flux. Although the atomic and ionised hydrogen components can be observed directly via their line emission, and are thus well constrained, this is not true for the molecular component. The molecular hydrogen column is assumed to be proportional to the CO emission line strength, as measured by the Columbia CO survey \\citep{Dam87}, for example, but the constant of proportionality (usually denoted by $X$) is unknown and one is forced to determine its value by fitting to the gamma-ray data. Although the uncertainty in the best-fit determination of $\\langle X\\rangle$ (averaged over the whole sky) is small, the systematic uncertainties are acknowledged to be much larger, ``at least 10\\%--15\\%'' \\citep{Hun97}. In turn, this estimate of the uncertainty is small in comparison with the differences among the various values of $X$ which have been reported in the literature (see the review by \\citet{Blo89}) and the likely range of variation in $X$ within the Galaxy \\citep{Hun97}. Although these are substantial uncertainties, molecular hydrogen contributes only a fraction of the total observed gamma-ray intensity roughly 20\\% of the local surface density of gas in the Galactic disk is thought to be in molecular form \\citep{Wou90} so the implied fractional uncertainty in the total Galactic emission is perhaps as small as 3\\%. Moreover, uncertainty in $X$ affects only the normalisation and not the spectral shape of the predicted emission from diffuse molecular hydrogen, so the observed high-energy excess cannot be explained in this way even if $X$ could assume an arbitrarily large value. The other main contributions to uncertainty in the diffuse model are associated with (1) unmodeled spatial variations in the cosmic-ray spectral energy density and (2) unmodeled spatial variations in the low-energy photon spectral energy density; the former affects all of the Galactic contributions to $I$, whereas the latter affects only the inverse Compton component. These uncertainties affect both the normalisation and the spectral shape of the predicted gamma-ray emission; however, the uncertainties are difficult to quantify. \\subsubsection{Constraints based on the gamma-ray data} For our purposes it is not actually necessary to quantify the uncertainties on the model input parameters; it suffices to use the discrepancy between model and data as a measure of the uncertainty in our understanding of the observed emission. In turn this measure determines the constraints which we can apply to any putative unmodeled emission, such as the contribution from dense gas which we are concerned with here. At photon energies $E>1\\;{\\rm GeV}$ the fractional discrepancy is roughly 60\\% \\citep{Hun97}, in the sense that the observed emission is 1.6 times larger than the model, and we henceforth adopt $0.6/1.6\\simeq40$\\% of the total observed intensity as our estimate of the unmodeled emission. Although this estimate is derived from data at high energies, the effects of the various contributing processes are all very widely spread, and {\\em the estimate therefore applies independent of photon energy.\\/} The constraints appropriate to high/low Galactic latitudes can now be re-evaluated. At high Galactic latitudes the observed intensity is $I\\simeq1.5\\times10^{-5}\\,{\\rm ph\\,cm^{-2}s^{-1}sr^{-1}}$ for $E\\ge100\\;{\\rm MeV}$ \\citep{Kni96}, implying that any unmodeled emission should be $\\la6\\times10^{-6}\\,{\\rm ph\\,cm^{-2}s^{-1}sr^{-1}}$ in this band. This result is actually slightly stricter than the criterion used by \\citet{Gil94} and thus leads us to tighten our high-latitude constraints, relative to those quoted in \\S3.2: the observed high-latitude gamma-ray intensity constrains the amount of low column-density gas to $\\la20$\\% of the total density of the Galactic dark halo, with this fraction rising to 100\\% for gas clouds of column density $\\Sigma\\ga200\\;{\\rm g\\,cm^{-2}}$. At low Galactic latitudes we can make use of the mean intensity of the inner Galactic disk, which has been accurately determined by \\citet{Hun97}. For example at 1~GeV the mean intensity ($|l|\\le60^\\circ$, $|b|\\le10^\\circ$) is $\\langle I\\rangle\\simeq 3\\times10^{-8}\\,{\\rm ph\\,cm^{-2}s^{-1}sr^{-1}MeV^{-1}}$, and our calculation of $\\langle Q\\rangle$ for this region yields (\\S2.3) $3.28\\times10^{-2}\\,{\\rm g\\,cm^{-2}}$, implying that the emissivity of the Galactic dark halo material must be, on average, ${\\cal E}\\le4.6\\times10^{-6}\\,{\\rm ph\\,s^{-1}g^{-1}MeV^{-1}}$. By comparison, the actual emissivity of low column-density gas is computed to be (\\S2.3, table 2) ${\\cal E}(1\\;{\\rm GeV})\\simeq1.4\\times10^{-5}\\,{\\rm ph\\,s^{-1}g^{-1}MeV^{-1}}$, implying that $\\la30$\\% of the Galaxy's dark halo may be comprised of low column density gas. For higher column densities the emissivity falls, and table 2 shows that for $\\Sigma=100\\,{\\rm g\\,cm^{-2}}$ the emissivity is only $5.5\\times10^{-6}\\,{\\rm ph\\,s^{-1}g^{-1}MeV^{-1}}$. The gamma-ray data on the inner Galactic disk thus indicate all of the Galaxy's dark halo to be made of dense clouds of column-density $\\Sigma\\ga100\\,{\\rm g\\,cm^{-2}}$. \\subsection{Comment on the gas content of the galactic disk} The constraint we have just given is based on the mean spectrum of the inner Galactic disk, in contrast to those given by \\citet{Sal96} who employed limits based on the angular structure of the observed gamma-ray intensity. Specifically, \\citet{Sal96} required that the putative contribution of emission from a baryonic component of the Galaxy's dark halo be less than that of the isotropic component of the observed intensity; this procedure seems to us to be less reliable than the procedure we have employed, for two reasons. First, even if the dark halo were spherically symmetric the emission from any baryonic component would not be, both because our point of observation is quite distant from the centre of the Galaxy and because the resulting gamma-ray emission is strongly dependent on the Galactic cosmic-ray distribution, which in turn is strongly concentrated in the disk of the Galaxy. Indeed the cosmic-ray distribution appears to correlate with the distribution of interstellar matter \\citep{Hun97}, thus complicating the interpretation of the observed gamma-ray intensity distribution. In particular this coupling leads to gamma-ray emission from a baryonic dark halo being correlated with the diffuse gas column density, even if the dense gas is uncorrelated with the diffuse gas. Second, on any given line of sight, such as the Ophiuchus field considered by \\citet{Sal96}, a highly structured dark matter halo might exhibit, by chance, a low dark matter column density. The chances of this are good if, as in the case of \\citet{Sal96}, the field is specifically chosen to have a low ``background'' intensity." }, "0403/astro-ph0403128.txt": { "abstract": "{intracluster medium --- active galaxies --- jets --- cooling flows --- feedback --- entropy} Jets and winds are significant channels for energy loss from accreting black holes. These outflows mechanically heat their surroundings, through shocks as well as gentler forms of heating. We discuss recent efforts to understand the nature and distribution of mechanical heating by central AGNs in clusters of galaxies, using numerical simulations and analytic models. Specifically, we will discuss whether the relatively gentle `effervescent heating' mechanism can compensate for radiative losses in the central regions of clusters, and account for the excess entropy observed at larger radii. ", "introduction": "Observational and theoretical lines of evidence suggest that accreting black holes in active galaxies release large amounts of kinetic energy to their environments. Observationally, a minority of active galaxies release the bulk of their power in the form of radio jets (Rees \\textit{et al.} 1982; Begelman \\textit{et al.} 1984). But more numerous radio quiet AGN may also generate much of their power in the form of fast winds, as indicated by recent spectral analyses of broad absorption line (BAL) QSOs in the optical/UV (Arav \\textit{et al.} 2001; de Kool \\textit{et al.} 2001) and X-ray (Reeves \\textit{et al.} 2003) bands. These studies suggest that the kinetic energy in BAL outflows --- which are thought to occur in most radio quiet QSOs --- is larger than previously thought and can approach the radiation output. Theoretical arguments also point to the importance of the kinetic energy channel. Numerical simulations suggest that accretion disks, which transfer angular momentum and dissipate binding energy via magnetorotational instability, may inevitably produce magnetically active coronae (Miller \\& Stone 2000). These are likely to generate outflows that are further boosted by centrifugal force (Blandford \\& Payne 1982). Unless radiation removes at least 2/3 of the liberated binding energy, very general theoretical arguments indicate that rotating accretion flows {\\it must} lose mass. The physical reason is that viscous stresses transport energy outward, in addition to angular momentum. If radiation does not remove most of this energy, then a substantial portion of the gas in the flow will gain enough energy to become unbound (Narayan \\& Yi 1995; Blandford \\& Begelman 1999, 2004). Blandford (this volume) discusses this effect and its consequences in more detail. While it may sometimes be possible to tune the system so that the gas circulates without escaping, any excess dissipation (i.e., increase of entropy) near a free surface of the flow will lead to outflow. There are several possible sources of such dissipation, including magnetic reconnection, shocks, radiative transport, and the magnetocentrifugal couple mentioned above. If radiative losses are very inefficient, outflows can remove all but a small fraction of the matter supplied at large radii. While it is probably safe to say that some substantial release of kinetic energy always accompanies accretion, we are not yet able to predict its magnitude, nor how it compares with radiative losses. During the growth of a supermassive black hole to mass $M_{BH}$, the integrated kinetic energy output could be comparable to the total release of binding energy, $\\sim 0.1 M_{BH} c^2$. It is unlikely to be much smaller than a tenth that amount. Even with a kinetic energy output of $\\sim 0.01 M_{BH} c^2$, the effects on the environment can be dramatic. At a kinetic energy conversion efficiency of $\\varepsilon_{KE} c^2$ per unit of accreted mass, an accreting black hole liberates $10^{19} (\\varepsilon_{KE}/0.01)$ ergs per gram. In a galactic bulge with a velocity dispersion of $200 \\sigma_{200}$ km s$^{-1}$, the accretion of one gram liberates enough energy to accelerate $2 \\times 10^4 (\\varepsilon_{KE}/0.01) \\sigma_{200}^2 $ grams to escape speed --- provided that most of the energy goes into acceleration. Since the typical ratio of black-hole mass to galactic bulge mass is $> 10^{-3}$ (H\\\"aring \\& Rix 2004), feedback from a supermassive black hole growing toward its final mass could easily exceed the binding energy of its host galaxy's bulge. Such feedback has been invoked in various models to explain the $M_{BH}-M_{bulge}$ and $M_{BH}-\\sigma$ correlations (e.g., Silk \\& Rees 1998; Blandford 1999; Fabian 1999). The effects of injecting a certain amount of kinetic energy into the environment of a supermassive black hole depend not only on the amount of energy injected, but also on the temporal and spatial characteristics of the injection process and the structure of the ambient medium. Explosive injection of a large amount of energy in a short time --- which might lead to intense, centrally concentrated heating at a shock --- will have very different consequences from gradual, intermittent, or spatially distributed injection, which would allow the surrounding medium to adjust. Whether mechanical heating is partially offset by radiative cooling is an important factor, as is the presence of absence of small-scale density inhomogeneities. The speed of a shock or sound wave propagating through a medium with a `cloudy' thermal phase structure will be highest in the phase with the lowest density (the intercloud medium). Dense regions will be overrun and left behind by the front, as first pointed out by McKee \\& Ostriker (1977) in connection with supernova blast waves propagating into the interstellar medium. Consequently, most of the energy goes into the gas which has the lowest density (and is the hottest) to begin with. The global geometric structure of the ambient gas is important as well. Since a wind or hot bubble emanating from an AGN will tend to follow the path of least resistance, a disk-like structure can lead to a `blowout' of hot gas along the axis, leaving the disk intact. Anisotropic injection of the kinetic energy, e.g., in a pair of jets or equatorial, can likewise affect certain regions while sparing others. Given these complications, it is especially desirable to find a `laboratory' where one can study the details of mechanical energy injections by active galaxies in a specific context. Clusters of galaxies can serve this role well. In the remainder of this paper we will discuss how recent X-ray observations of the intracluster medium (ICM) suggest a particular mode of heating by AGNs, which appears to be susceptible to theoretical modeling. While mechanical heating is likely to be the only important form of AGN feedback into the hot, highly ionized atmospheres of galaxy clusters, we stress that other forms of energy injection, particularly radiative heating, must occur as well. These effects are can be particularly important on smaller scales (i.e., the interstellar medium of the host galaxy) and in less highly ionized environments; they are discussed by Ostriker (this volume). ", "conclusions": "There is growing evidence, from observations as well as theory, that active galaxies can provide widespread mechanical heating of their environments. In clusters, this heating apparently occurs in a rather gentle fashion, and is driven by buoyancy. Simulations of buoyant plumes show significant and fast lateral spreading, generation of sound waves, cool rims of entrained gas surrounding the hot bubbles and a mismatch between X-ray and radio emission, resulting in `ghost cavities'. The evenness and spatial distribution of heating may be adequate to balance cooling globally, prevent cooling catastrophes, and thus quench so-called `cooling flows'. Energy spreading to larger radii during cluster assembly might be able to account for the observed entropy excesses in present-day clusters. While the initial results are promising, numerical simulations have a long way to go before they can adequately the represent the physics of AGN heating. Three-dimensional simulations are already being done, but their resolution needs to be increased in order to study the effects of mixing and thermal instability. Magnetic fields, which have played little role in models to date, may have important effects on the dynamics, transport properties (viscosity and thermal conduction), and radio emissivity of clusters. At the microphysical level, we need to understand why the hot (relativistic?) gas injected by active galaxies appears to mix relatively little with the ICM. Indeed, we cannot exclude the possibility that some of the injected `cosmic rays' do stream through the thermal background. If they couple effectively to the ICM via hydromagnetic waves, they will heat the gas in much the same fashion as expanding bubbles, as they traverse the pressure gradient (Loewenstein \\textit{et al.} 1991). Finally, we need to better understand what sets the efficiency of kinetic energy output from black hole accretion flows, the speed and degree of collimation of the output (winds vs. narrow jets), and feedback effects that couple the evolution of the ICM to the growth rate of the black hole. Whether (and how) such feedback fixes the $M_{BH}- M_{bulge}$ correlation by regulating the black hole mass, the galaxy mass, or both, remains to be seen." }, "0403/astro-ph0403406_arXiv.txt": { "abstract": "We apply the intrinsically symmetrical, decelerating relativistic jet model developed by Laing \\& Bridle for 3C\\,31 to deep, full-synthesis 8.4-GHz VLA imaging of the two low-luminosity radio galaxies B2\\,0326+39 and B2\\,1553+24. After some modifications to the functional forms used to describe the geometry, velocity, emissivity and magnetic-field structure, these models can accurately fit our data in both total intensity and linear polarization. We conclude that the jets in B2\\,0326+39 and B2\\,1553+24 are at angles of $64^\\circ\\pm5^\\circ$ and $7.7^\\circ\\pm1.3^\\circ$ to the line of sight, respectively. In both objects, we find that the jets decelerate from 0.7 -- 0.8$c$ to $<$0.2$c$ over a distance of approximately 10\\,kpc, although in B2\\,1553+24 this transition occurs much further from the nucleus than in B2\\,0326+39 or 3C\\,31. The longitudinal emissivity profiles can be divided into sections, each fit accurately by a power law; the indices of these power laws decrease with distance from the nucleus. B2\\,0326+39 also requires a discontinuity in emissivity to in order to fit a region with several bright knots of emission. In B2\\,1553+24, the sudden brightening of the jet can be explained by a combination of rapid expansion of the jet and a continuous variation of emissivity. The magnetic fields in both objects are dominated by the longitudinal component in the high-velocity regions close to the nucleus and by the toroidal component further out, but B2\\,0326+39 also has a significant radial component at large distances, whereas B2\\,1553+24 does not. Simple adiabatic models fail to fit the emissivity variations in the regions of high velocity but provide good descriptions of the emissivity after the jets have decelerated. Given the small angle to the line of sight inferred for B2\\,1553+24, there should be a significant population of similar sources at less extreme orientations. Such objects should have long ($\\ga$200\\,kpc), straight, faint jets and we show that their true sizes are likely to have been underestimated in existing images. ", "introduction": "\\label{intro} Evidence that the jets in low-luminosity, FR\\,I \\citep{FR74} radio galaxies are initially relativistic and decelerate on kpc scales has mounted in recent years. FR\\,I sources are thought to be the side-on counterparts of BL Lac objects, in which relativistic motion on parsec scales is well-established \\citep{UP95}. Superluminal motions have been seen on milliarcsecond scales in several FR\\,I jets \\citep{Giovannini01} and on arcsecond scales in M\\,87 \\citep{Biretta95}. In FR\\,I sources, the lobe containing the main (brighter) jet is less depolarized than the counter-jet lobe \\citep{Morganti97b}. This can be explained if the main jet points toward the observer, suggesting that the brightness asymmetry is caused by Doppler beaming \\citep{Laing88}. The asymmetry decreases with distance from the nucleus \\citep{Laing99}, implying that the jets must decelerate. Self-consistent models of the deceleration of relativistic jets by injection of matter lost from stars or entrained from the surrounding galactic atmosphere have been calculated by \\citet{Bicknell94}, \\citet{Komissarov94} and \\citet{Bowman96}. \\citet[][hereafter LB]{LB} fit VLA images of the radio jets in the nearby FR\\,I radio galaxy 3C\\,31 using a sophisticated model to reproduce the observed total and polarized emission within 30\\,arcsec of the nucleus, where the jets are straight. They parameterized the three-dimensional distributions of velocity, emissivity and magnetic-field structure, calculated the brightness at each point within the jets in Stokes $I$, $Q$ and $U$, accounting for the effects of relativistic aberration, and integrated along the line of sight to reproduce the expected distributions on the sky. They concluded that the jets in 3C\\,31 could be accurately modelled as intrinsically identical, axisymmetric, antiparallel, decelerating, relativistic flows, with locally random but anisotropic magnetic fields. Optimization of the model parameters placed tight constraints on the geometry, velocity, emissivity and field structure. \\citet{LB2} used this velocity field, together with estimates of the external pressure and density from {\\sl Chandra} observations \\citep{Hardcastle02} in a conservation-law analysis based on that of \\citet{Bicknell94}. They showed that there are self-consistent solutions for jet deceleration by injection of thermal matter and derived the variations of pressure, density, Mach number and entrainment rate along the jets. Finally, \\citet{Laing04} developed models of adiabatic, relativistic jets with velocity shear and applied them to 3C\\,31. They demonstrated that such models provide a reasonable description of the emissivity and magnetic-field variations at large distances from the nucleus but fail closer in, and inferred that significant reacceleration of relativistic particles is required precisely where X-ray synchrotron emission is observed \\citep{Hardcastle02}. In the present paper we apply a modified version of LB's model to the jets of two FR\\,I radio galaxies: B2\\,0326+39 and B2\\,1553+24. Our principal aim is identical to that of LB: to estimate the distributions of velocity, emissivity and magnetic-field structure without introducing preconceptions about the (poorly understood) internal physics. By studying different objects we hope to be able to improve the range of physical scales we are able to probe, to identify which intrinsic features are common to all FR\\,I jets and which vary from object to object and to assess the dependence of the jet structure on power and environmental conditions. Section \\ref{obs} presents our new, deep, full-synthesis VLA images. In Section \\ref{model}, we outline the model, emphasizing the parameterizations of the geometry, velocity, emissivity and magnetic field which differ from those used by LB. Section \\ref{results} compares our best-fitting models with the observed data in a variety of ways to show the features that we are able to reproduce as well at those we cannot. In Section \\ref{physical}, we present the velocity, emissivity and field structures of the best-fitting models. Section \\ref{discuss} investigates whether the magnetic-field structure and emissivity are consistent with the assumptions of flux freezing and pure adiabatic energy loss and examines the idea that the jets reaccelerate. We then consider the appearance of the model for B2\\,1553+24 at large angles to the line of sight and the implications for the detectability of side-on counterparts. Finally, we briefly compare the models for 3C\\,31, B2\\,0326+39 and B2\\,1553+24. Section \\ref{ssfw} summarizes our conclusions and outlines possible avenues for further work. We adopt a Hubble constant, $H_0$ = 70\\,$\\rm{km\\,s^{-1}\\,Mpc^{-1}}$ throughout and define spectral index $\\alpha$ in the sense $S(\\nu) \\propto \\nu^{-\\alpha}$. We use the notation $p = (Q^2+U^2)^{1/2}/I$ for the degree of linear polarization. ", "conclusions": "\\label{discuss} \\subsection{Adiabatic models} \\label{adiabatic} The simplest physical picture of the evolution of the emissivity along a jet assumes that the radiating particles only lose energy adiabatically (synchrotron and inverse-Compton losses being negligible by comparison), that there are no dissipative processes such as particle acceleration or field-line reconnection, and that the magnetic field is convected passively with the (laminar) flow. Analytical equations for the emissivity in this case were first derived by \\citet{Burch79}; these were generalized by \\citet{Baum97} to describe a relativistic, decelerating flow with purely perpendicular or parallel field. We use the latter expressions, modified to calculate the magnetic-field evolution self-consistently as in \\citet{Laing04}, to estimate the emissivity profiles expected for our models of jet shape and velocity. We refer to the fits derived in the previous sections and by LB as {\\em free models} in order to distinguish them from {\\em adiabatic models}, following the terminology of \\citet{Laing04}. \\subsubsection{Magnetic-field profiles} \\label{freeze} \\citet{Baum97} showed that the variations of the magnetic field components in the quasi-one-dimensional approximation are: \\begin{eqnarray*} B_r &\\propto& (x\\beta\\Gamma)^{-1} \\\\ B_t &\\propto& (x\\beta\\Gamma)^{-1} \\\\ B_l &\\propto& x^{-2} \\\\ \\end{eqnarray*} in the absence of shear ($x$ is again the jet radius). Given the field components at one point in the jet, we can then predict their evolution from the profiles of radius and velocity given in the previous section. The predicted evolution of the field components in B2\\,0326+39 and B2\\,1553+24 is shown by the dashed lines in Figs~\\ref{fig:0326bprofiles} and \\ref{fig:1553bprofiles}, respectively. The initial conditions have been chosen to match the free models at the distances from the nucleus given in the figure captions (note that the field is forced to be longitudinal at the nucleus). The qualitative evolution of the field components is consistent with the free models for both sources: a decrease in the longitudinal component is accompanied by an increase in the toroidal component. Rapid flaring and deceleration act in the same sense, leading to a large decrease in the relative fraction of longitudinal field, as observed. There are significant quantitative discrepancies, however, particularly in the regions where the jets decelerate. In both sources, the transition from longitudinal to transverse field is predicted to be less abrupt than is seen in the free models. This discrepancy is qualitatively consistent with the effects of velocity shear due to a transverse velocity gradient, which will act to increase the longitudinal component and to delay the onset of the longitudinal to transverse transition. This cannot be the whole story, however, as there are also large differences between the variations of the toroidal and radial components. The former increases, as expected, but the latter stays roughly constant: in the quasi-one-dimensional approximation, their ratio should remain constant. Similar problems occur in 3C\\,31 (LB; \\citealt{Laing04}). \\citet{KO89,KO90} showed that the variation of the rms toroidal and longitudinal field components with distance could be very different from that predicted by the simple adiabatic approximation if the flow is turbulent, as is widely believed (e.g.\\ \\citealt{Bic84,DeY96}). Given that the gross departures from the flux-freezing predictions occur where the jets are fast, or are decelerating rapidly, it seems likely that turbulence affects both the magnitude and the configuration of the field in these regions. \\subsubsection{Emissivity profiles} \\label{ademiss} The emissivity function $\\epsilon$ can be written in terms of the magnetic field $B$, both as defined in Section~\\ref{Emissivity}, as: \\begin{eqnarray*} \\epsilon \\propto (x^2\\beta\\Gamma)^{-(1+2\\alpha/3)}B^{1+\\alpha} \\end{eqnarray*} \\citep{Baum97,Laing04}. $B$ can be expressed in terms of the parallel-field fraction $f = \\langle B_l^2 \\rangle ^{1/2}/B$ and the radius $\\bar{x}$, velocity $\\bar{\\beta}$ and Lorentz factor $\\bar{\\Gamma}$ at some starting location using equation 8 of \\citet{Laing04}: \\begin{eqnarray*} B \\propto \\left[ f^2\\left(\\frac{\\bar{x}}{x}\\right)^4 + (1 - f^2)\\left(\\frac{\\bar{\\Gamma}\\bar{\\beta}\\bar{x}}{\\Gamma\\beta x}\\right)^2\\right ]^{1/2} \\end{eqnarray*} We can therefore predict the emissivity using our fitted jet width and velocity together with an estimate of the parallel-field fraction $f$. The resulting emissivity profiles for B2\\,0326+39 and B2\\,1553+24 are shown in Figs~\\ref{fig:0326profiles}(c) and \\ref{fig:1553profiles}(d), respectively. The solid lines show the emissivity profiles from our free model fits and the dashed lines the self-consistent adiabatic profiles, normalized to match at large $z$. In both objects the adiabatic models agree poorly with the free models where the jet velocities are high. The emissivity falls off much too rapidly in the adiabatic models, just as in 3C\\,31 (LB). This is not a surprise, for the following reasons: \\begin{enumerate} \\item Velocity shear (required by the free model) has been neglected. \\item The magnetic-field evolution is more complicated than expected from simple flux-freezing in an axisymmetric laminar-flow model, even if shear is included (Section~\\ref{freeze}) and turbulence may dominate the field evolution \\citep{KO89,KO90}. \\item Where the jets are fast, we see complex, small-scale, non-axisymmetric structures, indicating that the flow is not laminar. \\item In B2\\,1553+24, there is optical synchrotron emission from the base of the main jet \\citep{Parma03}, implying continuing particle acceleration. \\end{enumerate} Further from the nucleus, where the velocity has a low and roughly constant value, the adiabatic profile matches the free model very well in both objects. In B2\\,0326+39 this is within a region 4 to 12\\,kpc (Fig.~\\ref{fig:0326profiles}). In B2\\,1553+24, the region of low, constant velocity extends from 30 to 200\\,kpc and the adiabatic profile agrees very well with the free model over approximately 1.5 orders of magnitude in emissivity (Fig.~\\ref{fig:1553profiles}). This suggests that the outer jets in this source are modelled surprisingly well as constant-velocity, perpendicular-field, adiabatically-evolving flows. \\subsection{Sidedness profiles and reacceleration} \\label{acceleration} Our models, motivated by the gross features of the observed sidedness ratios, assume monotonic deceleration. There are theoretical reasons to expect jets to be reaccelerated by the pressure gradient of the external medium at large distances from the nucleus provided that the mass injection rate is not too large, for example if stellar mass loss dominates the mass injection \\citep{Komissarov94,Bowman96}. The most obvious effect of reacceleration is a small increase of sidedness ratio with distance from the nucleus. No such increase is obvious from the ridge-line profiles or images of sidedness ratio (Figs~\\ref{fig:0326ls} and \\ref{fig:1553ls}). We have therefore averaged the sidedness ratios in rings of constant distance from the nucleus in order to improve the signal-to-noise ratio (Fig.~\\ref{avside}). This reveals considerable sidedness structure in the low-velocity regions. Both sources show sidedness minima, at $\\approx$14\\,kpc from the nucleus in B2\\,0326+29 and at $\\approx$8\\,kpc in B2\\,1553+24. The minimum for B2\\,0326+39 occurs roughly where \\citet{Worrall00} inferred that the synchrotron minimum pressure in the jets \\citep{Bridle91} becomes less than the pressure of the external medium. Fig.~\\ref{avside} shows the average observed sidednesses as solid lines and our best-fitting models as dashed lines. The dotted lines show the profiles for models with the outer velocity parameters, $\\beta_0$, modified to fit the minimum and maximum sections of the observed profiles in the outer region. The velocity ranges are $\\beta_0 =$ 0.03 -- 0.25 for B2\\,0326+39 and 0.13 -- 0.19 for B2\\,1553+24. [Note that the model sidedness profiles increase slightly with distance from the nucleus because progressively larger areas of jet edge, which have lower sidedness ratios than the centres, fall below the intensity blanking threshold]. If the changes in sidedness result from acceleration, then they should be associated with variations in the degree of polarization. In B2\\,0326+39, the field structure can be roughly approximated by two-dimensional field sheets with equal radial and toroidal components (Section~\\ref{0326magnetic}). The expected changes in polarization are then straightforward to calculate \\citep{Laing80}. For an acceleration from $\\beta = 0.03$ to $\\beta = 0.25$ between 16 and 18\\,arcsec from the nucleus, as implied by a naive interpretation of Fig.~\\ref{avside}(a), $p$ should vary from 0.49 to 0.64 in the main jet and from 0.45 to 0.38 in the counter-jet. These changes are not observed (Figs~\\ref{fig:0326lp}e and \\ref{fig:0326avlp}). The velocity increase is far larger than is expected from the effects of any pressure gradient in the external medium and we also note that the fit of an accelerating adiabatic model to the emissivity profile would be significantly worse than that shown in Fig.~\\ref{fig:0326profiles}. In B2\\,1553+24, it is more difficult to exclude acceleration as the cause of the increase in sidedness ratio between 8 and 14\\,arcsec from the nucleus (Fig.~\\ref{avside}b): the predicted variations in degree of polarization are below the fluctuation level in Fig.~\\ref{fig:1553lp}(e) and we cannot average across the jet to reduce the noise; the velocity increase required ($\\beta_0 =$ 0.13 -- 0.19) is physically more reasonable than in B2\\,0326+39 and the adiabatic model fit to the emissivity is comparable in quality to that in Fig.~\\ref{fig:1553profiles}. We conclude that there is no compelling evidence in favour of reacceleration of the jets in either source, and that it is most unlikely to be the sole cause of the increase in sidedness ratio with distance from the nucleus in B2\\,0326+39. Nevertheless, it is expected theoretically, there are hints that it might occur, and it is consistent with the total intensity and polarization data in B2\\,1553+24. \\begin{figure*} \\includegraphics[width=17cm]{fig20.eps} \\caption{Sidedness profiles of: (a) B2\\,0326+39 at 0.5 arcsec resolution and (b) B2\\,1553+24 at 0.75 arcsec resolution, averaging over all points at the same distance from the nucleus. Solid line: data; dashed line: best fit model; dotted lines: models with $\\beta_0$ modified to fit the minimum and maximum sidednesses in the lower-velocity region, as described in the text. \\label{avside}} \\end{figure*} \\subsection{The appearance of B2\\,1553+24 at other angles to the line of sight} \\label{1553incl} We infer that the angle to the line of sight for B2\\,1553+24 is $\\approx$8$^\\circ$. Given that the B2 sample from which it is drawn is selected at the low frequency of 408\\,MHz, the usual assumption is that the emission is isotropic and therefore that the distribution of jet orientations for its members should be random. The source should therefore have many ($\\sim$100) counterparts of comparable total luminosity and larger $\\theta$, but the entire sample of B2 radio galaxies with jets defined in Tables\\,1 and 2 of \\citet{Parma87} has 43 members. A further concern is that the counterparts must be very large. The length of the main jet is at least 60\\,arcsec (Fig.~\\ref{fig:1553.montage}a), corresponding to 388\\,arcsec (340\\,kpc) at $\\theta = 60^\\circ$, the median angle to the line of sight. This is comparable in size with the longest jet in the sample, in NGC\\,315 \\citep{Willis81}, and far larger than the median ($\\approx$30\\,kpc; \\citealt{Parma87}). There is cause for suspicion unless: (a) the side-on counterparts of B2\\,1553+24 are not members of the B2 sample or (b) sources in that sample have linear sizes far larger than previously realised. In order to understand potential selection effects, we have computed the appearance of the model brightness distribution for B2\\,1553+24 at various angles to the line of sight. We show the results for the median angle to the line of sight for an isotropic sample ($\\theta = 60^\\circ$) and for a source in the plane of the sky ($\\theta = 90^\\circ$) in Fig.~\\ref{fig:1553los}. The area chosen for the plots corresponds to the modelled region for the $60^\\circ$ case. As well as being very long, the jets appear narrow, straight and (except for a small region around the nucleus) extremely faint. The total flux density from the inner 200\\,kpc of the model jets (measured along the jets and excluding the core) is constrained to be 44\\,mJy for the $\\theta = 7.7^\\circ$ model but is 27\\,mJy for $\\theta = 60^\\circ$. There are three important selection effects: \\begin{enumerate} \\item The extended flux of the jets shows significant Doppler boosting at small $\\theta$. Given that there is little lobe emission in B2\\,1553+24 \\citep{deRuiter93,NVSS}, the jets probably dominate the flux at low radio frequencies, so end-on sources of this type will be selected preferentially in the B2 survey. The integral source count $N(>S) \\propto S^{-3/2}$ for the luminosities and redshifts in question, so the size of the parent sample is effectively increased by a factor of $(44/27)^{3/2} \\approx 2$ if isotropic lobe emission is ignored. \\item Except for a small region around the nucleus, the model jets are faint. Further than 50\\,arcsec from the nucleus, the surface brightness is $\\la$0.16\\,mJy/beam area at 8.4\\,GHz with a beam of 2.75\\,arcsec FWHM. This scales to 0.47\\,mJy/beam area at 1.4\\,GHz for a spectral index $\\alpha = 0.6$, comparable with the detection threshold for typical VLA images of sources in the sample presented by \\citet{Fanti86,Fanti87}. In any case, emission on scales $\\ga$120\\,arcsec would not have been imaged reliably in existing VLA observations of the B2 sample (C configuration at 1.4\\,GHz; \\citealt{Fanti87}). The combination of these two effects makes it extremely unlikely that the outer jets would have been detected. \\item Sources of large angular size with flux densities close to the limit of the catalogue could have been missed by the original B2 survey, which measured peak rather than integrated flux density. We have estimated the magnitude of this effect by convolving the model brightness distribution for $\\theta = 60^\\circ$ with the beam of the B2 survey (3 $\\times$ 10\\,arcmin$^2$). If the short axis of the beam is parallel to the jet axis, the ratio of peak/total flux is 0.73; if the long axis of the beam is aligned, the ratio becomes 0.95. A few sources could therefore have been missed (primarily those orientated E-W). \\end{enumerate} We conclude that the side-on counterparts of B2\\,1553+24 could have escaped identification, either because they were missed in the original survey or because their angular sizes have been greatly underestimated. Morganti \\& Parma (private communication) and Ledlow \\& Owen (2004; in preparation) have made more sensitive observations of radio galaxies in the B2 sample using the WSRT and the VLA in D configuration and have shown that a significant fraction of them have much longer radio jets than have previously been reported, extending many 100's of kpc or even further. We suggest that these form part of the missing population. \\begin{figure*} \\includegraphics[height=17cm,angle=-90]{fig21.eps} \\caption{The brightness distribution of the model for the jets in B2\\,1553+24 at angles to the line of sight of: (a) $\\theta = 60^\\circ$ and (b) $\\theta = 90^\\circ$. The plots cover 394\\,arcsec (346\\,kpc) in projection on the sky. This corresponds to a distance from the nucleus along the jet axis of 200\\,kpc in both directions for $\\theta = 60^\\circ$, i.e.\\ to the size of the modelled region. The contour levels are 1, 2, 4, 8, 16, 32, 64, 128, 256, 512 $\\times$ 20$\\mu$Jy/beam and the model images have been convolved with a 2.75-arcsec Gaussian beam.\\label{fig:1553los}} \\end{figure*} \\subsection{Similarities and differences between B2\\,0326+39, B2\\,1553+24 and 3C\\,31} \\label{compare} In later papers, we will present a re-analysis of 3C\\,31 along with models of two further objects, all with the same functional descriptions of velocity, emissivity and magnetic field. This will allow a full comparison between them; here we make a few general observations about the similarities and differences between B2\\,0326+39, B2\\,1553+24 and 3C\\,31. Notable similarities are: \\begin{enumerate} \\item All three objects can be modelled successfully by a decelerating jet model. \\item The jets are all initially well collimated, then flare before recollimating and expanding conically. \\item Close to the nucleus the jet velocity on-axis is consistent with a value of $\\beta \\approx 0.8$ in all cases. \\item Deceleration occurs over a region of $\\approx$10\\,kpc to an outer velocity of $0.0 \\la \\beta \\la 0.2$. \\item The ratio of edge to on-axis velocity is consistent with a constant value of $\\approx$0.7 everywhere, although in some places this is poorly constrained. \\item The exponent of the emissivity index is similar in the outermost modelled regions of all the jets. \\item At the jet edge the emissivity falls to $\\approx$0.25 of its on-axis value. \\item The longitudinal field component dominates in the high-velocity regions close to the nucleus and the toroidal component in the outer parts, qualitatively but not quantitatively as expected from flux freezing. \\item The variation of the radial field component is more complex than predicted by simple flux-freezing models. This effect cannot simply be due to velocity shear in a laminar, axisymmetric flow. \\item Close to the nucleus (in the high-velocity and deceleration regions), the variation of emissivity with distance from the nucleus is far less rapid than predicted by the adiabatic approximation. \\item In the low-velocity outer regions, the emissivity variation is reasonably well fitted by a quasi-one-dimensional adiabatic model. \\end{enumerate} The three objects were expected to have a wide range of angles to the line of sight; this is confirmed by our modelling. In addition, there are some clear differences: \\begin{enumerate} \\item The conical outer region of 3C\\,31 has a large opening angle and is centred on the nucleus. The outer regions of the other two objects are closer to cylindrical, with very small opening angles and vertices far behind the nucleus. [Note that 3C\\,31 does recollimate outside the modelled region (LB)]. \\item The velocity in the outer region of 3C\\,31 was modelled by LB as decreasing monotonically with distance from the nucleus (as required by the variation of sidedness ratio). In the B2 sources, it has an approximately constant asymptotic value. \\item The close connection between the jet geometry and the forms of the velocity and emissivity profiles inferred for 3C\\,31 is not general. \\item The velocity in B2\\,1553+24 remains at $\\beta \\approx 0.8$ until $\\approx$20\\,kpc from the nucleus (cf.\\ $\\approx$2\\,kpc in the other two objects). \\item There is no need for a discontinuity in emissivity in B2\\,1553+24: the brightening of the jet is consistent with expansion at constant emissivity. \\item Although the biggest single field component at large distances is always toroidal, there are significant differences in the details of the field-component evolution: 3C\\,31 has a mixture of longitudinal and toroidal components; B2\\,0326+39 has toroidal and radial components and B2\\,1553+24 has almost pure toroidal field. \\item In 3C\\,31, the radial field component in the flaring region increases towards the edge of the jet. There is no evidence for this effect in the B2 sources, although it cannot be excluded in B2\\,0326+39, where the signal-to-noise ratio is low. \\item The adiabatic approximation describes the emissivity evolution in B2,1553+24 over a much larger fraction of the jets than in the other two sources. This may be because it applies most accurately in the low-velocity outer region, which is relatively longer in this source. \\end{enumerate}" }, "0403/astro-ph0403630_arXiv.txt": { "abstract": "A Monte-Carlo calculation of the atmospheric neutrino fluxes~\\cite{BGS,AGLS} has been extended to take account of the three-dimensional (3D) nature of the problem, including the bending of secondary particles in the geomagnetic field. Emphasis has been placed on minimizing the approximations when introducing the 3D considerations. In this paper, we describe the techniques used and quantify the effects of the small approximations which remain. We compare 3D and 1D calculations using the same physics input in order to evaluate the conditions under which the 3D calculation is required and when the considerably simpler 1D calculation is adequate. We find that the 1D and 3D results are essentially identical for $E_\\nu>5$~GeV except for small effects in the azimuthal distributions due to bending of the secondary muon by the geomagnetic field during their propagation in the atmosphere. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403540_arXiv.txt": { "abstract": "{ Possible forsterite nanoparticle model of the Extended Red Emission(ERE) is proposed on the basis of photo-luminescence of forsterite after gamma-ray and neutron irradiation. Forsterite exhibits interesting thermoluminescence spectrum similar to ERE of Red Rectangle after irradiation in low temperature. It is shown that the forsterite after thermoluminescence is over exhibits photo-luminescence(PL) when Ultraviolet ray is irradiated. The structure of PL spectrum is almost similar to that of thermoluminescence. In order to explain small variations of the peak position of wavelength of ERE spectrum, possible nanoparticle model of forsterite is investigated. Our model is consistent to the ISO observation data in near and middle infrared region, which suggest the existence of forsterite. ", "introduction": "Extended red emission (ERE) is a broad emission band with a peak wavelength between 600 and 850 nm, and with a width between 60 and 120 nm seen in many dusty astrophysical objects such as reflection nebulae, planetary nebulae, HII regions, halos of galaxies, and even in the Diffuse Interstellar Medium. The observation of ERE in the Diffuse Interstellar Medium shows that ERE is a general phenomenon. Though the carrier for ERE is not yet clear, the proposed carriers are hydrogenate amorphous carbon (HAC), quenched carbonaceous composite (QCC), $\\rm C_{60}$, carbon nanoparticles, polycyclic aromatic hydrocarbons (PAHs), and silicon nanoparticles, and most of them appear to be unable to explain the observed ERE spectra (see Ledoux et al. 2001; Witt et al. 1998, for a summary). We have suggested that thermoluminescence spectra of forsterite after $\\gamma$-ray irradiation are very similar to ERE of Red Rectangle (Koike K. et al., 2002, here after KK), and have discussed such possibility that thermoluminescence is related to the changes of property of interstellar and circumstellar matter by various irradiation in that space. It is, however, not so plausible that the irradiation energy is sufficient to explain such emissions. Recently, we have found that the forsterite after thermoluminescence is over exhibits photo-luminescence(PL) when ultraviolet ray (UV) is irradiated. This fact seems to suggest a possible realistic mechanism of ERE. This paper is concerning to this problem. \\begin{figure*} \\includegraphics[width=17cm,height=8cm]{Forste_UV-F0.eps} \\caption{Photoluminescence spectra of forsterite $(\\rm Mg_2SiO_4)$ single crystals under UV irradiation. The sample is firstly irradiated by $(\\gamma)$-ray in liquid nitrogen, and warmed till about 500K. After thermoluminescence has been over, it is irradiated by UV of a mercury lamp at room temperature.} \\label{Forste_UV_R} \\end{figure*} Interstellar and circumstellar matter is irradiated by high energy electromagnetic and cosmic ray particles such as $\\gamma$ rays, neutrons, protons and heavy-ions etc. These irradiation will cause some changes on properties such as optical ones of these materials. Especially, it is known that extremely large fluxes of neutrons and $\\gamma$-rays have been emitted during super-nova explosions. Furthermore, interstellar and circumstellar space is typically at extremely low temperature and is always irradiated by electromagnetic radiation and by cosmic ray particles for long time-scale. The effect of this radiation will accumulate in the low temperature environment. In the circumstellar region of both young and evolved stars as well as in the solar system, forsterite and enstatite have been found (Waters et al., 1998ab ; Malfait et al., 1998; Wooden et al., 1999) by the Infrared Space Observatory (ISO) (Kessler et al., 1996). Carbonates such as dolomite $\\rm CaMg(CO_{3})_{2}$, breunnerite $\\rm Mg(Fe,Mn)(CO_{3})_{2}$, calcite $\\rm CaCO_3$, and $\\rm Mg$,$\\rm Ca$-bearing siderite $\\rm FeCO_3$ were found in CI chondrite (Endress, Zinner and Bischoff, 1996). Among these carbonates, Ca-bearing minerals such as dolomite and calcite were also detected in dust shells around evolved stars by ISO (Kemper et al., 2002). Especially, it should be noted that the broad emission feature responsible for extended red emission (ERE) appears at about the 600 -- 900 nm region in many reflection nebulae, and among reflection nebulae the Red Rectangle nebula shows stronger intensity by one order(Witt and Boroson, 1990). In the Red Rectangle nebula, both PAH- and crystalline silicates (forsterite and enstatite)-features were observed by ISO (Waters et al., 1998b). We have suggested, on the basis of the measurement of thermoluminescence for irradiated silicates and carbonates, our thermoluminescence spectra of forsterite at 645--655 nm is very similar to the ERE of the Red Rectangle (KK, 2002). In this paper, we will investigate the changes of properties of silicates by irradiation in the context of astrophysics. Especially, it is emphasized that the forsterite after thermoluminescence is over exhibits photo-luminescence(PL) when ultraviolet ray (UV) is irradiated. The structure of PL spectrum is almost similar to that of thermoluminescence. Further, possible forsterite nanoparticle model of ERE is also discussed on the basis of this fact. \\begin{figure} \\centering \\includegraphics[width=7.5cm]{Forste_UV-Red-F-2.eps} \\caption{Comparison with Observation Data of Red Rectangle Spectrum (From Witt A.N. and Boroson T.A. 1990) and Photoluminescence spectra of forsterite under UV irradiation (Fig.1) } \\label{Forste_UV-Red} \\end{figure} ", "conclusions": "It should be noted that interstellar and circumstellar space is typically at extremely low temperature and is always irradiated by both electromagnetic radiation and by cosmic ray particles over cosmological timescales. Furthermore, it is well known that extremely large fluxes of neutrons and gamma-rays are emitted during supernova explosions. We have ever suggested that thermoluminescence spectra of forsterite after $\\gamma$-ray irradiation are very similar to ERE of Red Rectangle. However, the energy efficiency of thermoluminescence seems to be not sufficient to explain ERE. In this paper, we have discussed on the fact that the photo-luminescence of irradiated forsterite exhibits similar spectrum to that of thermoluminescence. It should also be remembered that forsterite and enstatite have been found by many ISO observations in many oxygen-rich young and evolved stars. Is the irradiated forsterite possible to be really one of the carrier of ERE? It is possible so far as its temperature is below about 1000K, because it is known in many insulators the effects of irradiation are almost maintained. However, the irradiation effects on insulators are known to fade out by annealing it to high temperature above this degree. In such case, can forsterite be a carrier of ERE? In semiconductor silicon, it is well known that the existence of a small amount of impurity caused by a certain kind of minor elements such as As, P, B, Ga or In etc cause new energy levels and forms new type of semiconductors. Though little is known about forsterite, it seems to be possible that a certain kind of impurities cause some semiconductor-like structure of forsterite. Investigation of such possibility is further problem. In this context, it should be noted that the melting point of silicon is $1410^{\\circ}C$, while that of forsterite is very high, and is $1890^{\\circ}C$. It is known that the existence of forsterite is observed in addition to ISO spectrum, in meteorites, while silicon itself is not found in both of them. Thus, together with the investigation of formation condition of silicates in the previous section, the silicon nanoparticle model seems to be highly hypothetical one in the present stage. We can get easily high-quality crystalline silicon because it is the most fundamental semiconductor, where the solid state properties is very well studied together with the problem of impurity. As for the forsterite, it is very difficult to get the high-quality one. Then, we have used the synthesized one by CZ method in laboratory. For the impurity component in our forsterite sample, Takei and Kobayashi (\\cite{takei}) reported previously that spectrographic analysis shows that the forsterite sample is pure. However, neutron activation analysis reveals that it contains an infinitesimal quantity of Ir at about 16--18 wt ppm. This level of Ir impurity is below the detection limit of spectrographic analysis. We have also measured our sample using radio activation analysis, and confirmed that our sample is almost pure; that is, other elements except for $\\rm Mg, Si, O$ and an infinitesimal quantity of Ir are not detected. In order to confirm the reappearance of thermoluminescence of forsterite in other sample, we have get forsterite powder for china and porcelain from a pottery \"Marusu\" (Japan), where details of the method of creation of it is not open. We have investigated thermoluminescence of the \"Marusu\" forsterite, and get almost the same result as our synthesized one." }, "0403/astro-ph0403589_arXiv.txt": { "abstract": "We have modeled the unusual orbital light curve of V1494 Aquilae (Nova Aquilae 1999 No.2) and found that such an unusual orbital light curve can be reproduced when there exist two-armed, spiral shocks on the accretion disk. V1494 Aql is a fast classical nova and found to be an eclipsing system with the orbital period of 0.1346138 days in the late phase of the nova outburst. Its orbital light curve shows a small bump at orbital phase 0.2, a small dip at 0.3, sometimes a small bump at 0.4, and a large bump at $0.6-0.7$ outside eclipse. Such a double- or triple-wave pattern outside eclipse has never been observed even though overall patterns look like some supersoft X-ray sources or eclipsing polars. We have calculated orbital light curves including the irradiation effects of the accretion disk and the companion by the hot white dwarf. These unusual patterns can be reproduced when we assume two-armed spiral shocks on the accretion disk. Especially, triple-wave patterns are naturally obtained. This result strongly suggests the existence of two-armed spiral shocks on the accretion disk in the late phase of the nova outburst. ", "introduction": "Angular momentum transport plays an essential role in accretion disks of cataclysmic variables. Two mechanisms have been proposed so far: one is the turbulent viscosity as adopted in the standard accretion disk model of \\citet{sha73} and the other is the direct dissipation by tidal spiral shocks on the accretion disk as first demonstrated by \\citet*{saw86}. The turbulent viscosity is a local physical process while the tidal spiral shocks have a global structure on the accretion disk. Therefore, we have a chance to observe a global spiral shock structure when they play an essential role in the angular momentum transport of the accretion disk. Such an observational evidence first came from the Doppler maps of the dwarf nova IP Peg outburst \\citep*[e.g.,][]{ste97}. We have long believed that tidal spiral structures can be detected even in orbital light curves of cataclysmic variables if the spiral patterns are prominent. At last we find such an evidence of spiral patterns on the accretion disk from the orbital light curves of V1494~Aquilae. V1494~Aquilae (Nova Aquilae 1999 No.2) was discovered by A. Pereira on 1999 December 1.785 UT at $m_V \\sim 6.0$ \\citep{per99}. It reached the visual maximum brightness of $m_V \\sim 4.0$ on Dec 3.4 UT ($t_0=$JD 2,451,515.9$\\pm 0.1$) and decayed at the rate of $t_2= 6.6 \\pm 0.5$ days and $t_3= 16 \\pm 0.5$ days \\citep{kis00}. Early phase spectra were taken by \\citet{fuj99} and \\citet{aya99}, which show P-Cyg profiles of hydrogen Balmer lines with a blueshifted component of 1020 km~s$^{-1}$ and 1200 km~s$^{-1}$, respectively. Thus, V1494~Aql has been classified as a fast nova. \\citet{kat04t} have summarized overall development of the V1494~Aql light curve until autumn of 2003 (for about 3 years). Detailed spectral features have been reported by \\citet{iij03}. A short-term periodic modulation was first reported by \\citet{nov00}, who found 0.03 mag variations with a period of 0.0627 days from their 2000 June observation. \\citet{ret00} analyzed 31 nights of CCD photometry during 2000 June-August and suggested a periodicity of 0.13467(2) days. The full amplitude of the variation in the $R$-band increased from 0.03 mag in June to 0.07 mag in August. Their folded light curve shows a double-wave structure with a shallower dip at phase 0.5, with about half the amplitude of the main periodicity. \\citet{bos01} reported, based on unfiltered and $R$-band CCD photometry obtained on 12 nights during 2001 June-July, a robust change in the shape and amplitude of the 0.13467 days period. It had an eclipse shape with depth about 0.5 mag. A second shallow eclipse (about 0.1 mag deep) at phase 0.5 can be seen. \\citet{bar03} refined, based on $V$-band photometry during June and September of 2002, the orbital period of 0.1346141(5) days. They pointed out that the orbital light curve is quite unusual in the sense that it is not similar to ordinary cataclysmic variables with a hot spot on the accretion disk. \\citet*{pav03} made a multicolor photometry and concluded that the eclipse depth is deeper in longer wave lengths (deeper in $I$-band than in $V$-band). \\citet{pav03} further discussed that such an eclipsing characteristic (deeper in longer wave lengths) cannot be explained by the accretion disk or the irradiation effect of the companion because these two light sources have the opposite nature (deeper in short wave lengths). Thus, they suggested, as a model of V1494~Aql orbital light curve, a self eclipsing accretion column in a magnetic polar. \\citet{kat04t} provided a time-development of the orbital light curves among 2001 November-December, 2002 August, and 2003 June. It sometimes shows a triple-wave structure as well as a double-wave structure. Kato et al. suggested that some structures (probably on the accretion disk) fixed in the binary rotational frame are responsible for the orbital light curve because of the stability of out-of-eclipse light curve patterns. Very recently, \\citet*{kis04} pointed out that a close companion to V1494~Aql (located $1\\farcs4$ southwest) is brighter than V1494~Aql itself in the very late phase of the outburst. They corrected the brightness of the orbital light curve and found that the depth of the eclipse is about twice deeper in magnitudes than before correction. The depth of the primary eclipse of V1494~Aql has become deeper as the nova decayed. Such a feature was also observed in the recurrent nova CI~Aql 2000 outburst \\citep[e.g.,][]{mat01}, in which the irradiation of the accretion disk plays an essential role \\citep{hac03ka}. Therefore, we expect the accretion disk in V1494~Aql is also responsible for the unusual wavy structure of the orbital light curve. In this Letter, we model the orbital light curve. Almost the same orbital light curve model as in the supersoft X-ray sources is adopted \\citep*[e.g.,][]{sch97, hac03kb, hac03kc}, which includes the irradiation effects of the accretion disk and of the companion. In addition, we assume two-armed spiral structures on the accretion disk to reproduce a triple-wave structure outside eclipse. In \\S 2, our numerical model for V1494~Aql is briefly introduced and summarized. The numerical results are given in \\S 3. Discussion follows in \\S4. \\placefigure{v1494_figure_late_i785} \\placetable{system_parameters_V1494_Aql} ", "conclusions": "The orbital light curve varies from night to night as shown in Figure 3 of \\citet{kat04t}. Such a scatter is also clearly shown in the present Figures 2 and 3 for Kiss et al.'s (2004) data. To reproduce this variation, we have changed parameter $\\xi_1$ and $\\xi_2$ independently (see Fig. \\ref{vmag_wd1000m03_quiescence_kiss_2}). The variation appears in orbital phase of $0.1-0.3$ and the upper and lower bounds for these variations can be well reproduced by the models of $\\xi_1=0.26$, $\\xi_2=0.44$ for the upper bound, and $\\xi_1=0.44$, $\\xi_2=0.24$ for the lower bound. Here, $\\xi_1$ and $\\xi_2$ represent the height of the spiral arm in the rear and front side on the disk of Figure \\ref{v1494_figure_late_i785}, respectively. The other parameters are the same as those in Figures \\ref{v1494_figure_late_i785} and \\ref{vmag_wd1000m03_quiescence_kiss}. It is remarkable that the peak at orbital phase $\\sim 0.65$ is rather stable and never disappears. This is consistent with the robustness of the period for the out-of-eclipse shape as emphasized by \\citet{kat04t}. \\citet{pav03} discussed the main light source for the wavy structures of the out-of-eclipse orbital light curve. They argued, as a light source, (1) accretion disk, (2) ellipsoidal shape of the companion, (3) reflection (irradiation) effect of the companion by the hot white dwarf, (4) intermediate polar activity, and (5) polar activity self-eclipsed by the accretion column. They rejected possibilities (1)-(4) and suggested a viable model of polar activity. Their reasons rejecting possibilities (2) and (4) are reasonable but not for possibilities (1) and (3) as discussed below. \\citet{pav03} rejected the possibility of accretion disk by pointing out two reasons: (i) the maximum duration of the eclipse of the disk limited by the Roche lobe size and it cannot exceed a quarter of the orbital phase while the overall duration of the eclipse is as large as 0.45 of the orbital period. (ii) The peak of light emission of the accretion disk lies in the blue region of the spectrum, so that the amplitude of the eclipse should increase as the wavelength becomes shorter (``as the wave length is increased'' in their text is mistaken). However, they observed the opposite trend: the amplitude in the red region is much more deeper than in the visual region. As for the first conjecture (i), based on the accretion disk model together with the irradiation effects of the disk and the companion, we have already constructed the orbital light curves that match well the observational data. Second, we should be careful with the above statement (ii) because the depth of the primary eclipse varies from night to night as seen in Figure 3 of \\citet{kat04t} and its depth changes as large as $\\Delta R_c \\sim 0.4$ between different periods of the observations as shown in Figures 2 and 3. Pavlenko et al.'s (2003) $R$ and $I$ light curves are not simultaneous ones but taken in different periods ($R$ is earlier than $I$). The $V$ light curve is reconstructed from Barsukova \\& Goranskii's (2003) data. They also rejected possibility (3) based on the same statement as (ii) discussed above. However, their multi-color results may simply suggest the fact that the eclipse depth varies from period (night) to period (night). Therefore, we cannot conclude statement (ii) only from the different depths at the different periods. For the assumed WD luminosity of $L_{\\rm WD}= 3,000~L_\\sun$, the distance modulus is obtained to be $(m-M)_R = 12.05$ and the corresponding distance is estimated to be $d \\sim 1.4$ kpc (see Table \\ref{system_parameters_V1494_Aql}). However, this does not mean the real distance to V1494~Aql because we do not know the real luminosity of $L_{\\rm WD}$ at the time of the two observations. For instance, if we adopt $L_{\\rm WD}= 750~L_\\sun$, the shape of the orbital light curve hardly changes but the distance modulus becomes $(m-M)_R = 11.45$ and $d \\sim 1$ kpc." }, "0403/astro-ph0403010_arXiv.txt": { "abstract": "Refraction of wave propagation in a corotating pulsar magnetospheric plasma is considered as a possible interpretation for observed asymmetric pulse profiles with multiple components. The pulsar radio emission produced inside the magnetosphere propagates outward through the rotating magnetosphere, subject to refraction by the intervening plasma that is spatially inhomogeneous. Both effects of a relativistic distribution of the plasma and rotation on wave propagation are considered. It is shown that refraction coupled with rotation can produce asymmetric conal structures of the profile. The differential refraction due to the rotation can cause the conal structures to skew toward the rotation direction and lead to asymmetry in relative intensities between the leading and trailing components. Both of these features are potentially observable. ", "introduction": "Pulsar radio emission is thought to originate in the region deep inside the pulsar magnetosphere populated with a relativistic electron-positron pair plasma (e.g. Blaskiewicz, Cordes, \\& Wasserman 1991; Melrose 2000). The radio waves propagate through the magnetosphere, subject to reflection and refraction (e.g. Barnard \\& Arons 1986; Petrova 2000; Fussell, Luo \\& Melrose 2003, and hereafter FLM) or absorption (Blandford \\& Scharlemann 1976; Lyubarskii \\& Petrova 1998; Luo \\& Melrose 2001; FLM). Since these propagation effects can give rise to observable features in the pulse profile and polarization, study of them can provide insight to the radiation processes inside the magnetosphere. In the polar cap model, relativistic electron positron plasmas are produced in the cascade above the polar cap (PC) as a result of rotation-induced particle acceleration (Sturrock 1971; Ruderman \\& Sutherland 1975; Arons \\& Scharlemann 1979; Daugherty \\& Harding 1982; Zhang \\& Harding 2000; Hibschman \\& Arons 2001). Primary electrons (positrons) are accelerated to ultra-high energies by a rotation-induced parallel electric field, emitting high energy photons through curvature radiation or inverse Compton scattering. High energy photons decay into electron/positron pairs in the strong pulsar magnetic field, forming an outflowing relativistic pair plasma, referred to as the pulsar plasma, which has a very broad distribution in parallel momenta. Although the mechanism for the radio emission is not well understood, it is generally believed that the emission is produced from collective plasma processes in the pulsar plasma, through either maser emission or plasma instabilities (e.g. Melrose 2000). Regardless of the specific emission mechanism, the radiation must be produced directly on or converted to modes that can escape the magnetosphere to reach the observer. In the pulsar plasma, there are two escape modes, the X mode with polarization perpendicular to the local plane of the magnetic field and wave vector, and the LO mode with polarization in this plane (e.g. Kennett, Melrose \\& Luo 2000, hereafter KML, and references therein). Observations of jumps between two orthogonal polarizations in the position angle of the radiation support the hypothesis that the radio emission propagates through the magnetosphere in these two orthogonal modes (Stinebring et al. 1984; McKinnon \\& Stinebring 2000). Therefore we only consider propagation of these two modes. A pulsar plasma is spatially inhomogeneous due to the radial and transverse dependence of the density in the open field line region. Wave propagation in such an inhomogeneous medium can be affected by refraction and reflection, which ultimately determines the emergent pulse profile. Because the dipole field, $B$, decreases in the radial direction as $B\\propto1/R^3$, where $R$ is the radial distance to the star's center, the density of the plasma that streams along the field lines must have the same radial dependence, $N\\propto 1/R^3$, corresponding to the longitudinal characteristic length scale for the density variation, $L\\sim R$, considerably larger than the star's radius. Variation in the plasma density in the direction perpendicular to the field lines is mainly due to the electron/positron pair cascade being nonuniform across the PC, which gives rise to a much smaller (than $R$) transverse length scale of order the radius of the open field line cone. Although the inhomogeneities can be on a still much smaller scale, possibly as a result of nonstationary acceleration or highly localized pair cascades, we only discuss the case that both radial and transverse scales are much larger than the relevant wavelength and that the refraction is the dominant effect that changes the ray path. Wave propagation can be then treated in the geometric optics approximation in which two orthogonal modes propagate independently of each other (Barnard \\& Arons 1986). The effect of refraction on wave propagation in the pulsar magnetosphere was discussed in the cold plasma approximation by several authors (e.g. Barnard \\& Arons 1986; Petrova 2000; Weltevrede et al. 2003). Numerical study of electron/positron pair cascades above the PC shows that the distribution of a pulsar plasma has a relativistic spread in the plasma rest frame (e.g. Daugherty \\& Harding 1982; Zhang \\& Harding 2000; Hibschman \\& Arons 2001; Arendt \\& Eilek 2002). The relativistic distribution strongly modifies the wave properties (KML) and therefore a self-consistent model for wave propagation needs to include the relativistic effect of the plasma. Past work on refraction of wave propagation generally ignores rotation and the ray path initially in the field line direction would remain two dimensional, confined to the plane of the magnetic field lines. The resultant pulse profile has the same symmetry as the intensity distribution at the emission origin (e.g. Petrova 2000). Observations often show pulse profiles with asymmetric multiple components, i.e. there is asymmetry in intensity as well as location in pulse longitude of conal components (Lyne \\& Manchester 1988; Gangadhara \\& Gupta 1998; Gupta \\& Gangadhara 2003). It is believed that distortion of the profile is due to aberration (Blaskiewicz, Cordes \\& Wasserman 1991; Hirano \\& Gwinn 2001; Gupta \\& Gangadhara 2003) or absorption (Luo \\& Melrose 2001; FLM). Wave propagation in rotating magnetospheres was recently discussed in FLM with emphasis on the cyclotron absorption. Cyclotron absorption including rotation can lead to differential absorption producing asymmetric pulse profiles. In their discussion, FLM assumed the resonance region to be at a substantial fraction of the light cylinder (the radius at which the corotation speed equals the light speed, $c$) where refraction can be ignored. The purpose of this paper is to explore the effect of refraction on wave propagation in the relativistic pulsar plasma that corotates with the star and the implication of such effect for interpretation of pulse profiles. We consider effects of both a relativistic distribution of the plasma and rotation. Due to rotation rays that originate from the leading and trailing components are subject to different refraction. It is suggested that such differential refraction can significantly distort the pulse profile. Strong refraction occurs in the region with a large density gradient, which is referred to as the refraction region and is located well below the cyclotron resonance region (FLM). To concentrate on the effect of refraction, one assumes the strong magnetic field approximation, in which the X mode is not affected by the medium and propagates approximately in a straight line through the magnetosphere. The LO mode is strongly refracted and is discussed here. Following a similar procedure to that described in FLM, the ray path in the rotating medium is evaluated numerically in the geometric optics formalism. The emergent pulse profile is obtained at radial distances beyond which the refraction is no longer effective. In Sec. 2, the wave dispersion of relativistic pulsar plasma is summarized with emphasis on the LO mode in the strong magnetic field approximation; The formalism of ray tracing in the rotating magnetosphere is described in Sec. 3. The ray path is obtained by numerically solving the ray equations including the relativistic distribution and rotation. The result is applied to interpretation of asymmetry of pulse profiles (Sec. 4). ", "conclusions": "The effect of refraction on wave propagation in rotating magnetospheres is considered including both effects of relativistic distribution and rotation. Since the relativistic plasma dispersion function is peaked at around $z\\approx1$ and remains small otherwise, inclusion of the intrinsic relativistic effect tends to suppress of refraction at frequencies $\\omega\\geq\\omega_c$. Refraction is sensitive to the plasma density profile in the transverse direction and causes rays to focus and bifurcate, which is qualitatively similar to the result from the nonrotation case (e.g. Petrova 2000). The focusing effect tends to produce a core component even when the emission region has a conal geometry. The bifurcation leads to split of the emission cone into two or more nested cones, giving rise to profiles with multiple components. One of the distinct features of the model discussed here is the prediction of asymmetry in the pulse profile. The conal components are skewed toward the rotation direction, which is qualitatively in agreement with recent observations (e.g. Gangadhara \\& Gupta 2001). The predicted profiles also show asymmetry in relative intensities between the leading and trailing components, which are common in observed pulse profiles. It is suggested that similar asymmetry seen in observations can be produced by combination of aberration and the differential refraction due to rotation. The latter is the dominant cause for the asymmetry if the radiation is produced at low altitudes. Two other modes that are not discussed here include the X mode and low-frequency Alfv\\'en mode. In general, pulsar plasma is gyrotropic due to that the electron and positron distributions are not identical. Although the X mode is not purely transverse, with a small longitudinal component of along the magnetic field, one expects refraction of the X mode to be weaker for the LO mode. Inclusion of the X mode would lead to pulse profiles of the two modes that are displaced with one mode dominating any particular part of the pulse longitude. The observational implication of such a displaced profile needs to be explored. The low-frequency Alfv\\'en mode can be refracted in the low-frequency regime. However, the wave becomes subluminal as it propagates to underdense regions and is subject to strong damping. Pair production above the PC can be oscillatory over the characteristic time scale about $\\Delta t_0=h_0/c$, where $h_0$ is the typical length of polar gap acceleration. The pulsar plasma formed from the cascade may consist of many outflowing clouds (e.g. Usov 1987; Asseo \\& Melikidze 1998). For $h_0\\sim R_*$, we have $\\Delta t_0\\sim 10^{-4}\\,\\rm s$. Since this time can be shorter than the ray propagation time, further work on the wave propagation including the time-dependence of the medium is needed." }, "0403/astro-ph0403226_arXiv.txt": { "abstract": "Present in over $45\\%$ of local spirals, boxy and peanut-shaped bulges are generally interpreted as edge-on bars and may represent a key phase in bar evolution. Aiming to test such claims, the kinematic properties of self-consistent 3D N-body simulations of bar-unstable disks are studied. Using Gauss-Hermite polynomials to describe the major-axis stellar kinematics, a number of characteristic bar signatures are identified in edge-on disks: 1) a major-axis light profile with a quasi-exponential central peak and a plateau at moderate radii (Freeman Type~II profile); 2) a ``double-hump'' rotation curve; 3) a sometime flat central velocity dispersion peak with a plateau at moderate radii and occasional local central minimum and secondary peak; 4) an $h_3-V$ correlation over the projected bar length. All those kinematic features are spatially correlated and can easily be understood from the orbital structure of barred disks. They thus provide a reliable and easy-to-use tool to identify edge-on bars. Interestingly, they are all produced without dissipation and are increasingly realized to be common in spirals, lending support to bar-driven evolution scenarios for bulge formation. So called ``figure-of-eight'' position-velocity diagrams are never observed, as expected for realistic orbital configurations. Although not uniquely related to triaxiality, line-of-sight velocity distributions with a high velocity tail (i.e.\\ an $h_3-V$ correlation) appear as particularly promising tracers of bars. The stellar kinematic features identified grow in strength as the bar evolves and vary little for small inclination variations. Many can be used to trace the bar length. Comparisons with observations are encouraging and support the view that boxy and peanut-shaped bulges are simply thick bars viewed edge-on. ", "introduction": "} Three-dimensional (3D) N-body simulations of bar-unstable disks have consistently shown that, soon after a bar forms, it buckles and settles with an increased thickness and velocity dispersion, appearing boxy or peanut-shaped (B/PS) when viewed edge-on \\citep[e.g.][]{cs81,cdfp90,rsjk91}. This is particularly important in view of the large number of vertically extended structures commonly refered to as B/PS bulges in edge-on spirals. The recent survey of \\citet*{ldp00a} reveals that at least 45\\% of bulges across all morphological types are B/PS \\citep[but see also the earlier studies of][]{j86,sa87,s87}, suggesting that a significant fraction of bulges may form through dynamical instabilities in disks rather than dissipational collapse \\citep*[e.g.][]{els62} or accretion of smaller systems \\citep[e.g.][]{sz78}. Verifying such claims is hard, as reliably identifying bars in edge-on systems is difficult. A plateau in the major-axis light profile of an edge-on spiral has long been claimed to indicate the presence of a bar \\citep*[e.g.][]{cc87,hw89,ldp00b}, but axisymmetric features could equally give rise to it and end-on bars would likely remain undetected. A kinematic identification is clearly called for, analogous to the use of longitude-velocity diagrams in the Galaxy (\\citealt{p75,ml86,bgsbu91}; etc). This was first proposed and used in the context of external bulges by \\citet{km95}, while \\citeauthor*{ba99} (\\citeyear{ba99}, hereafter \\citeauthor{ba99}) and \\citeauthor*{ab99} (\\citeyear{ab99}, hereafter \\citeauthor{ab99}) refined those kinematic bar diagnostics using, respectively, periodic orbit calculations and hydrodynamical simulations. These were then applied to relatively large samples of galaxies by \\citet{mk99} and \\citet{bf99}, who successfully showed a close relationship between B/PS structures and bars. Using the ionized-gas kinematics, these works were however only able to probe the galactic potentials in the equatorial plane and were restricted to intermediate and late-type spirals. Using N-body simulations, we develop in this paper fully self-consistent stellar kinematic bar diagnostics for edge-on systems. Those diagnostics are most easily applicable to gas-poor early-type spirals (where the bulges are large and contain little dust, perfect for photometric studies), and they allow to probe galactic potentials out of the disk plane (to test the barred nature of B/PS structures at large galactic heights). The diagnostics rely exclusively on readily observable quantities and are thus not meant as a study of barred N-body models. They have already been successfully applied by \\citet{cb04} to the complete sample of \\citet{bf99}, and they are being applied at different heights to a subset of the galaxies by Zamojski et al.\\ (2005, in preparation). \\citet{baabdf05} and \\citet*{aab05} present $K$-band observations of the same sample, quantifying the B/PS structure. The goal of all those studies is to study the vertical structure of bars, and ultimately to clarify their relationship to bulges in the context of bar-driven evolution scenarios \\citep[e.g.][]{pn90,fb93,fb95,a03a}. We review existing bar diagnostics for edge-on disks in \\S~\\ref{sec:past_diagn} and describe the N-body simulations used throughout this paper in \\S~\\ref{sec:nbody}. The basic stellar kinematic bar diagnostics are presented in \\S~\\ref{sec:bar_diagn} along with their viewing angle dependence. \\S~\\ref{sec:time_evol} and \\S~\\ref{sec:incl_dep} describe, respectively, the time evolution of the diagnostics (as the bar evolves) and their inclination dependence (for galaxies close to but not perfectly edge-on). We discuss the uniqueness of the bar diagnostics and compare them to existing observations in \\S~\\ref{sec:discussion}, and our results are summarized in \\S~\\ref{sec:conclusions}. ", "conclusions": "} To constrain bar-driven evolution scenarios for the formation of bulges, and more specifically to provide tools to probe the nature of boxy and peanut-shaped (B/PS) bulges, we have used self-consistent 3D N-body simulations of bar-unstable disks to study the kinematic signatures of edge-on bars. Quantifying the major-axis stellar kinematics with Gauss-Hermite polynomials, a number of features can be used as bar diagnostics: 1) a steep quasi-exponential central light profile with a shoulder or plateau at moderate radii; 2) a double-hump rotation curve, possibly showing a local maximum and minimum at moderate radii; 3) a sometime broad central velocity dispersion peak with a shoulder or plateau (and possibly a secondary maximum) at moderate radii; 4) line-of-sight velocity distributions (LOSVDs) with a high-velocity tail (i.e.\\ an $h_3-V$ correlation) over the projected bar length. A local central $\\sigma$ minimum can also be present for strong bars seen approximately end-on. The positions and lengths of those features are further correlated, providing an easy and reliable tool to identify bars in edge-on disks. We note that despite previous claims, a ``figure-of-eight'' is never seen in stellar position-velocity diagrams (PVDs), as expected from realistic models and orbital configurations. Interestingly, while some of the kinematic bar features can individually be created by axisymmetric density distributions, the $h_3-V$ correlation appears to be a particularly reliable tracer of triaxiality (although it is not uniquely related to it). A number of those kinematic bar signatures can also be used as proxies for the bar length, and thus as an indirect measure of the bar pattern speed. The sharpness of the kinematic features generally decreases for weaker bar and/or increasing viewing angle (from end-on to side-on), introducing some degeneracy between the two but ensuring complementary with morphological bar signatures (strongest for side-on bars). All characteristic features of the kinematic profiles are established the moment the bar forms and grow stronger with time, as the bar lengthens and strengthens. They vary little for small inclination variations but noticeable differences appear for $i\\lesssim80\\degr$. Existing data and the recent work of \\citet{cb04} show that most B/PS bulges indeed show the kinematic signatures identified here, lending support to evolution scenarios where those bulges are formed through vertical disk instabilities. Detailed work on the morphological and photometric properties of both simulated and observed B/PS bulges are ongoing \\citep{baabdf05,aab05} and should provide more tests of the consistency of these scenarios. Preliminary results support our current conclusions. The current simulations also systematically produce Freeman Type~II (truncated) surface brightness profiles with approximately exponential peaks, as is observed in most bulges (at least late-types). The local central $\\sigma$ minimum present in the strongest bars (and produced without dissipation) is also increasingly thought to be common in spirals." }, "0403/astro-ph0403156_arXiv.txt": { "abstract": "{We present infrared H- and K-band spectra of a companion candidate $3^{\\prime \\prime}$ north of the young star GSC~08047-00232, a probable member of the nearby young Horologium association. From previously obtained JHK-band colors and the magnitude difference between primary and companion candidate, the latter could well be substellar (Neuh\\\"auser et al. 2003) with the spectral type being roughly M7-L9 from the JHK colors (Chauvin et al. 2003). With the H- and K-band spectra now obtained with ISAAC at the VLT, the spectral type of the companion candidate is found to be M6-9.5. Assuming the same age and distance as for the primary star ($\\sim 35$ Myrs, 50 to 85 pc), this yields a mass of $\\sim 25$ Jupiter masses for the companion, hence indeed substellar. After TWA-5 B and HR 7329 B, this is the third brown dwarf companion around a nearby ($\\le 100$~pc) young ($\\le 100$~Myrs) star. A total of three confirmed brown dwarf companions (any mass, separation $\\ge 50$ AU) around 79 stars surveyed in three young nearby associations corresponds to a frequency of $6 \\pm 4~\\%$ (with a correction for missing companions which are almost on the same line-of-sight as the primary star instead of being separated well), consistent with the expectation, if binaries have the same mass function as field stars. Hence, it seems that there is no brown dwarf desert at wide separations. ", "introduction": "Brown dwarfs (BDs) are objects with mass below the hydrogen burning mass limit, i.e. below $\\sim 0.078$~M$_{\\odot}$ (e.g. Burrows et al. 1997). It is not yet clear, how BDs form. According to one of the suggested scenarios, accreting companions get ejected as BD embryos by encounters inside a forming multiple system (e.g. Reipurth \\& Clarke 2001). In such a case, BDs may be different from stars regarding multiplicity, kinematics, and disk properties, so that such a scenario can be tested observationaly, e.g. by comparing the secular evolution of star-star binaries with star-BD binaries: If single BDs became single by the ejection from a multiple system, then the fraction of star-BD binaries should decrease faster with time than star-star binaries. A possible secular evolution in stellar binaries has been studied by Bouvier et al. (2001) and Patience et al. (2002) with as yet inconclusive results. In order to compare such results with star-BD binary fractions, one would first need significant statistics. Hence, several groups search for BDs as companions to stars. As far as young stars ($\\le 100$ Myrs) in nearby associations ($\\le 100$ pc) are concerned, only two brown dwarfs have been confirmed so far as companions by both common proper motion as well as spectroscopy, namely TWA-5~B (Lowrance et al. 1999, Neuh\\\"auser et al. 2000) in the TW Hya association (TWA, Webb et al. 1999) and HR~7329~B (Lowrance et al. 2000, Guenther et al. 2001) in the Horologium association (HorA, Torres et al. 2000) and/or $\\beta$ Pic moving group (Zuckerman \\& Webb 2000). In Sect. 2, we present previous and new imaging observation and investigate the astrometry, i.e. whether the pair is co-moving. Our spectroscopic observations, the data reduction, and the final H- and K-band spectra are shown in Sect. 3. We obtain a mass estimate for the companion and interprete the results in Sect. 4. A discussion in terms of BD companion frequency is given in the last section. ", "conclusions": "" }, "0403/gr-qc0403111_arXiv.txt": { "abstract": "We discuss the Sagnac effect in standard Minkowski coordinates and with an alternative synchronization convention. We find that both approaches lead to the same result without any contradictions. When applying standard coordinates to the complete rim of the rotating disk, a time-lag has to be taken into account which accounts for the global anisotropy. We propose a closed Minkowski space-time as an exact equivalent to the rim of a disk, both in the rotating and non-rotating case. In this way the Sagnac effect can be explained as being purely topological, neglecting the radial acceleration altogether. This proves that the rim of the disk can be treated as an inertial system. In the same context we discuss the twin paradox and find that the standard scenario is equivalent to an unaccelerated version in a closed space-time. The closed topology leads to preferred frame effects which can be detected only globally. The relation of synchronization conventions to the measurement of lengths is discussed in the context of Ehrenfest's paradox. This leads to a confirmation of the classical arguments by Ehrenfest and Einstein. ", "introduction": "The interpretation of the Sagnac effect is a longstanding problem in the theory of special relativity. Although different approaches of explanation agree on the observable effects, which are in turn consistent with experiments, the interpretation in the context of special relativity is still a matter of debate, as the continuously high publication rate on the subject shows. The problem is closely related with time measurements in moving reference frames and especially with the synchronization of clocks at different positions in rotating systems. This paper is not intended to be a review of previous work on the subject but instead tries to derive the theory and conventions from first principles in so far as it seems appropriate in the given context. It is meant to be more or less self-contained, so that some overlap with previous articles cannot be avoided. We do not claim a full axiomatic foundation of special relativity, like the one presented in the book of \\citetp{reichenbach24},\\nocite{reichenbach69} of course. For an overview of the Sagnac effect and closely related subjects, including experiments, theoretical analysis and philosophical interpretation, we refer to the articles of \\citetp{post67}, \\citetp{hasselbach93}, \\citet{stedman97} and especially the extensive discussion of synchronization issues by \\citet{anderson98}\\phantom{.} These reviews also include extended lists of references for further reading. Very recently, \\citet{rrf04} compiled a book about ``Relativity in Rotating Frames'' containing a number of articles about the Sagnac effect, its interpretation and related issues. The article by \\citet{rizzi04} themselves comprises another review of the subject. The outline of this paper is as follows. First we will briefly describe the problem the Sagnac effect poses for special relativity. After this introduction we will describe the principles of relativity. The Sagnac effect will be described using the standard conventions, finding that this is possible without contradictions. This proves that special relativity can be used also on the circumference of a rotating disk. However, we will see that the standard coordinates are not valid globally in this case, which leads us to the discussion of a more general approach. We will derive general coordinates and transformations which are still equivalent and compatible with standard relativity although they use a different convention. We will argue that the synchronization can be chosen arbitrarily and show that in certain situations the choice of a non-standard convention can be more convenient, although not necessary. We will discuss why the standard coordinates are not valid globally in the situation of a rotating disk, and show that the problem can be formulated without explicitly considering the acceleration. Instead, the closed topology of the rim of a rotating disk has to be taken into account. Locally (which will be defined), the rim is indistinguishable from Minkowski space. Only if our measurement process somehow encloses the complete circle, do we notice the rotation and are able to detect any non-trivial effects. This in a natural way leads us to the central point of this paper, which is the topological interpretation. We will construct a (spatially) closed Minkowski space-time which will be revealed to be exactly equivalent to the rim of a rotating disk. With this model at hand, we can easily avoid some of the difficulties present in previous discussions of the Sagnac effect. This proves that acceleration is not the main reason for the problems of interpretation. Special relativity is valid, and the rim of the rotating disk can even be seen as an inertial frame, as long as the radial direction is not probed by the experiment. After this discussion, we will turn to the twin paradox which is shown to be closely related to the Sagnac effect, especially when it is discussed on the rotating disk or equivalently in closed space-time. As before, we will propose a scenario which avoids any acceleration effects while it still leads to exactly the same physical effects as the standard situation. In this way the twin paradox is interpreted in terms of space-time topology as well. Finally, we will discuss measurements of lengths, which are related with the synchronization problem. We will find that, although the synchronization of clocks is a matter of convention, spatial lengths of objects at rest have a meaning as invariant intervals in space-time only if these are measured along lines of standard simultaneity. The measurement of lengths of moving objects, on the other hand, is more a matter of discussion. The result will generally depend on the details of the experiment which implicitly applies a synchronization convention in order to define the lengths. This will lead to our interpretation of the Ehrenfest paradox which agrees with classical approaches but disagrees with several recent publications on the subject. In all our presentation we will avoid any sophisticated mathematical methods in order not to hide the physical content behind the formalism. We put the emphasis on a self-consistent and logical derivation of the arguments, without hiding any implicit assumptions. Although we tried to include references to most of the relevant publications, we do not claim completeness in the discussion of previous work. ", "conclusions": "All the discussions in this paper have in common the importance of clear definitions of the concepts used. Once \\emph{time} and \\emph{length/space} are defined unambiguously, the measurements in all the (real and thought) experiments can be translated relatively easily into their corresponding space-time concepts. The rim of a rotating disk can be treated as an inertial system without any contradiction, as long as the radial coordinate is not probed. Unless the disk is surrounded, the situation is exactly equivalent to that of linear motion as discussed in standard relativity textbooks. When the global structure, i.e.\\ the topology, of the rotating circle becomes relevant, one has to ensure that the locally valid coordinates and space-time descriptions really match or ``mesh'' globally. On the rotating disk, we found that local Minkowski coordinates can not be extended globally, which gives the explanation for the Sagnac effect. In order to overcome any lack of confidence in our notion that the acceleration is not important for the discussion of effects on the rim of the rotating disk, we showed the equivalence with the situation of a topologically closed Minkowski space-time, which can physically not be distinguished from the rim of a rotating disk. We have to keep in mind that the standard notion of special relativity using Lorentz transformations leads to coordinates which are valid locally. In standard open space-time situations, these coordinates can be extended to form globally valid systems. Periodic boundary conditions or closed space-time topology, on the other hand, will restrict the allowed range of these coordinates. Additional discontinuities or time-lags can appear when matching the local coordinates globally. This leads to preferred frame effects which are purely global and of topological nature. Although most fundamental physics is usually formulated in terms of partial differential equations, we have to keep in mind that such a local description can naturally not explain the world completely. Boundary conditions and thus the topology of space-time can play an equally important role. \\newcommand{\\mnras}{Mon.\\ Not.\\ R.\\ Astron.\\ Soc.} \\newcommand{\\pz}{Phys.\\ Z.} \\newcommand{\\fp}{Found.\\ Phys.} \\newcommand{\\fpl}{Found.\\ Phys.\\ Lett.} \\newcommand{\\physrep}{Phys.\\ Rep.} \\newcommand{\\grg}{Gen.\\ Relativ.\\ Gravit.} \\newcommand{\\ajp}{Am.\\ J.\\ Phys.} \\newcommand{\\ejp}{Europ.\\ J.\\ Phys.} \\newcommand{\\cqg}{Class.\\ Quant.\\ Gravit.} \\newcommand{\\grqc}{ArXiv Gen.\\ Relat.\\ Quant.\\ Cosmol.\\ e-print} \\newcommand{\\physics}{ArXiv Phys.\\ e-print} \\newcommand{\\ncim}{Nuovo Cimento} \\newcommand{\\rpp}{Rep.\\ Prog.\\ Phys.} \\newcommand{\\ppsa}{Proc.\\ Phys.\\ Soc.\\ London, Sect.~A} \\newcommand{\\comptrend}{C.\\ R.\\ Acad.\\ Sci.} \\newcommand{\\jdp}{J.\\ Phys.\\ (Paris)} \\newcommand{\\annphys}{Ann.\\ Phys.\\ (Leipzig)} \\let\\cite=\\ocite" }, "0403/astro-ph0403464.txt": { "abstract": "{One of the goals of the ground-based support program for the \\corot\\ and \\romer\\ satellite missions is to characterize suitable target stars for the part of the missions dedicated to asteroseismology. We present the detailed abundance analysis of nine of the potential \\corot\\ main targets using the semi-automatic software \\vwa. For two additional \\corot\\ targets we could not perform the analysis due to the high rotational velocity of these stars. For five stars with low rotational velocity we have also performed abundance analysis by a classical equivalent width method in order to test the reliability of the \\vwa\\ software. The agreement between the different methods is good. We find that it is necessary to measure abundances extracted from each line relative to the abundances found from a spectrum of the Sun in order to remove systematic errors. We have constrained the global atmospheric parameters \\teff, \\logg, and [Fe/H] to within $70-100$~K, $0.1-0.2$~dex, and 0.1~dex for five stars which are slow rotators ($\\vsini < 15$~\\kms). For most of the stars stars we find good agreement with the parameters found from line depth ratios, \\halpha\\ lines, Str\\\"omgren indices, previous spectroscopic studies, and also \\logg\\ determined from the \\hipparcos\\ parallaxes. For the fast rotators ($\\vsini > 60$~\\kms) it is not possible to constrain the atmospheric parameters. ", "introduction": "\\corot\\ (COnvection, ROtation, and planetary Transits) is a small space mission, dedicated to asteroseismology and the search for exo-planets (Baglin \\etal~2001). Among the targets of the asteroseismology part of the mission, a few bright stars will be monitored continuously over a period of 150 days. These bright targets will be chosen from a list of a dozen F \\& G-type stars, located in the continuous viewing zone of the instrument. The final choice of targets needs to be made early in the project, as it will impact on some technical aspects of the mission. A precise and reliable knowledge of the candidate targets is required, in order to optimize this final choice. Among the information which needs to be gathered on the candidate targets, fundamental parameters like effective temperature, surface gravity, and metallicity will play a major role in the selection of targets for \\corot. Projected rotation velocities, as well as detailed abundances of the main chemical elements will also be taken into account in the selection process. {\\ bf Thus, the aim of this study is to obtain improved values for the fundamental parameters and abundances of individual elements for the \\corot\\ main targets.} This information on the targets will subsequently be used for the selected stars, in conjunction with asteroseismological data obtained by \\corot, to provide additional constraints for the modelling of the interior and the atmospheres of these stars. % Above: Contrib from CATALA % Below: Hans' org. text In \\sect~\\ref{observations} we summarize the spectroscopic observations and data reduction, in \\sect~\\ref{sec:fundamental} we discuss the determination of the fundamental parameters from spectroscopy, photometry and \\hipparcos\\ parallaxes and we summarize previous spectroscopic studies of the target stars. In \\sect~\\ref{sec:methods} we describe the three different methods we have used for abundance analysis. In \\sect~\\ref{constrain} we discuss how we constrain the fundamental atmospheric parameters using abundance results for a grid of models. In \\sect~\\ref{sec:abundances} we discuss the abundances we have determined. Lastly, we give our conclusions in \\sect~\\ref{conclusions}. \\begin{table} \\centering \\caption[]{Log of the observations for the spectra of the proposed \\corot\\ targets we have analysed. The signal-to-noise ratio in the last column is calculated around 6\\,500 \\AA\\ in bins of 3~\\kms. \\label{tab:log}} \\begin{tabular}{r|rc|r|r} \\hline HD & \\multicolumn{1}{c}{Date} & UT start & $t_{\\rm exp}$ & S/N \\\\ \\hline 43318 & 19-Jan-98 & 22:38 & 900 & 120 \\\\ 43587 & 14-Jan-98 & 22:32 & 1800 & 250 \\\\ 45067 & 15-Jan-98 & 22:20 & 1200 & 260 \\\\ 49434 & 17-Jan-98 & 22:33 & 600 & 160 \\\\ 49933 & 21-Jan-98 & 22:17 & 1200 & 210 \\\\ 55057 & 15-Jan-98 & 23:10 & 900 & 270 \\\\ 57006 & 10-Dec-00 & 00:41 & 1800 & 250 \\\\ 171834 & 05-Sep-98 & 19:00 & 1800 & 300 \\\\ 184663 & 18-Jun-00 & 01:30 & 1000 & 170 \\\\ 46304 & 17-Jan-98 & 22:05 & 600 & 170 \\\\ 174866 & 04-Jul-01 & 00:39 & 1500 & 190 \\\\ \\hline \\end{tabular} \\end{table} \\begin{table} \\centering %\\footnotesize \\caption{Str\\\"omgren photometric indices of the \\corot\\ main targets taken from Hauck \\& Mermilliod (1998). HD~46304 and HD~174866 are shown separately: abundance analysis has not been made for these two stars due to their high \\vsini.\\label{tab:stromgren_corot}} \\begin{tabular}{r|ccccc} \\hline \\multicolumn{1}{c|}{HD} & $V$ & $b-y$ & $m_1$ & $c_1$ & ${\\rm H}_\\beta$ \\\\ \\hline 43318 & 5.65 & 0.322 & 0.154 & 0.446 & 2.644 \\\\ 43587 & 5.70 & 0.384 & 0.187 & 0.349 & 2.601 \\\\ 45067 & 5.87 & 0.361 & 0.168 & 0.396 & 2.611 \\\\ 49434 & 5.74 & 0.178 & 0.178 & 0.717 & 2.755 \\\\ 49933 & 5.76 & 0.270 & 0.127 & 0.460 & 2.662 \\\\ 55057 & 5.45 & 0.185 & 0.184 & 0.876 & 2.757 \\\\ 57006 & 5.91 & 0.340 & 0.168 & 0.472 & 2.625 \\\\ 171384 & 5.45 & 0.254 & 0.145 & 0.560 & 2.682 \\\\ 184663 & 6.37 & 0.275 & 0.149 & 0.476 & 2.665 \\\\ \\hline 46304 & 5.60 & 0.158 & 0.175 & 0.816 & 2.767 \\\\ 174866 & 6.33 & 0.122 & 0.178 & 0.960 & 2.822 \\\\ \\hline \\end{tabular} \\end{table} ", "conclusions": "} \\begin{itemize} \\item We have performed a detailed abundance analysis of nine potential \\corot\\ main targets. The accuracy of the abundances of the main elements is of the order 0.10~dex when including possible errors on microturbulence and inadequacies of the applied 1D LTE atmospheric models. \\item We have compared three different methods for the analysis which show very good agreement. The discrepancies are due to the different parameters we have used for the stellar atmosphere models. \\item We have found no evidence for chemically peculiar stars. \\item We have constrained \\teff, \\logg, and metallicity to within $70-100$~K, $0.1-0.2$~dex, and 0.1~dex for the five slow rotators in our sample. For the four moderate rotators we cannot constrain the fundamental parameters very well, \\ie\\ \\teff, \\logg, and [Fe/H] to within $200$~K, $0.5$~dex, and 0.15~dex. \\item For most of the stars our results for the fundamental parameters agree with the initial estimates from Str\\\"omgren photometry, line depth ratios, and \\halpha\\ lines, and the \\hipp\\ parallaxes. For HD~43318 we have found a somewhat lower \\teff\\ and \\logg. For HD~49933 we find a \\teff\\ about 200~K hotter than previous studies. For the fast rotators HD~171834 and HD~184663 we find a quite high \\logg\\ value compared to other methods, but the uncertainty on our estimate is large (0.5 dex). \\item For the two \\corot\\ targets HD~46304 and HD~174866 abundance analyses were not possible due to the very high \\vsini. \\end{itemize} Suggestions for future studies of the \\corot\\ targets: \\begin{itemize} \\item From the comparison of two independent analyses (method A and C; \\cf~\\ref{sec_comp_meth}) we have found evidence that a careful (re-)normalization of the spectra may be very important but this must be investigated further. \\item To probe the interior of the stars with the asteroseismic data from \\corot\\ we need to know the abundances of elements which affect the evolution of the stars, namely C, N, and O. Thus, spectra that cover the infrared regions with good C, N, and O lines must be obtained. % \\item To observe spectra with higher resolution could improve % the abundance estimates and hence the ``fine-tuning'' of the atmospheric % parameters. % This will give more precise EQW -- but this is not the % % main uncertainty here !! Models mv. are far worse error % % sources. \\item Recent 3D atmospheric models should also be used in the analysis to explore the importance of the models. % We have shown that the \\vwa\\ software % can potentially be used for larger samples of stars. % We note that a future paper in this series % will present similar results for the proposed \\romer\\ targets. \\end{itemize} We finally note that the next paper in this series will give the results of the abundance analysis for some recently proposed \\corot\\ primary targets, the possible secondary \\corot\\ targets, as well as the proposed targets for the \\romer\\ mission. %==================================================================" }, "0403/astro-ph0403360_arXiv.txt": { "abstract": "We have identified two quiescent GRBs (bursts having two or more widely-separated distinct emission episodes) in which the post-quiescent emission exhibits distinctly different characteristics than the pre-quiescent emission. In these two cases (BATSE GRBs 960530 and 980125), the second emission episode has a longer lag, a smoother morphology, and softer spectral evolution than the first episode. Although the pre-quiescent emission satisfies the standard internal shock paradigm, we demonstrate that the post-quiescent emission is more consistent with external shocks. We infer that some observed soft, faint, long-lag GRBs are external shocks in which the internal shock signature is not observed. We further note that the peak luminosity ratio between quiescent episodes is not in agreement with the ratio predicated by the lag vs. peak luminosity relationship. We briefly discuss these observations in terms of current collapsar jet models. ", "introduction": "There is strong evidence that the majority of gamma-ray burst (GRB) pulses result from internal shocks in relativistic winds \\citep{sar97a,kss97,dm98,rrf00,np02}. However, a number of bursts exhibit a soft component indicative of external shocks that could be interpreted as onset of afterglow \\citep{gib99,bur99,gib02}. This emission begins preferentially towards the end of the burst or even after the GRB has ended (as suggested by the BeppoSAX WFC and NFI observations of serveral x-ray afterglows \\citep{cos00}), but can also appear at a time early enough to overlap the short-timescale emission (as observed in GRB 980923 \\citep{gib99}). In addition, extended soft $\\gamma-$ray emission has been observed by co-adding the fluxes of many BATSE bursts; this might also be an indicator of the same external shock phenomenon \\citep{con02}. ``Quiescent'' GRBs may provide new clues to the origins and physics of GRB shocks. Quiescent bursts are GRBs that release their gamma-ray energy in more than one distinct emission episode \\citep{rm01,rmr01}; e.g. they have at least one extended period during which emission is absent. A rigorous definition of quiescent time can be problematic due to the diverse nature or complexity of GRB temporal structures. \\citet{rm01} introduced a quiescent GRB definition for a sample of bright BATSE bursts that used a sliding temporal window with a width of $0.05 \\times {\\rm T90}$ on summed, background-subtracted, BATSE 4-channel 64-ms data. These authors chose GRBs which had at least one time period for which the count rate dropped below $2\\sigma$ of the summed background rate. This criterion was applied to only the brightest, longest bursts in the BATSE 4B Catalog \\citep{pac99}. It is possible that this definition is too permissive, since it might allow for inclusion of some GRBs that do not have distinct emission episodes. For example, some bright bursts that do not satisfy this criterion (because the first episode's decay and the second episode's rise overlap) might be classified as quiescent if they are observed at lower signal-to-noise, as might some bursts with faint, long, underlying pulses. Because the existing quiescent definition might be too permissive, we have tentatively further required quiescent GRBs to be those in which the emission drops to the background for an extended period of time that is greater than or equal to the duration of the previous emission episode. This ensures that the quiescent time is greater than the interpulse duration. In their study of very bright quiescent GRBs, \\citet{rm01} operated under the assumption that the post-quiescent episodes are due to internal shocks. We offer the hypothesis that quiescent GRBs potentially provide a laboratory in which {\\it external} shock signatures indicative of the onset of afterglow may be isolated from the rest of the burst. We have thus far found two GRBS that support this hypothesis. ", "conclusions": "At least two quiescent GRBs exhibit multiple lags. The short-lag component is consistent with internal shocks while an external shock explanation is preferred for the smooth long-lag component. The long-lag component is broad and faint, and is possibly present but unrecognizable in many bursts. The apodization of \\citep{nor00} implicitly recognizes its existence in order to remove it. It is possible that many or all of the long-lag bursts in the sample of \\citet{nor02} exhibit only the smooth and broad long-lag component. Because the short and long lags found in these quiescent GRBs are so different, we have assumed that there are two distinct GRB pulse types. This hypothesis needs to be verified. The best way to approach this problem is via statistical and/or data mining studies of GRB pulse characteristics. It is important to determine the existence of one continuous pulse distribution or poossibly distinctly different distributions. Detailed analyses of BATSE and $Swift$ GRB data are planned to corroborate this work. The lag vs. peak luminosity relation apparently requires re-calibration. It is not clear that long-lag pulses obey the same relation as short-lag pulses, or if they even obey a lag vs. peak luminosity relation. Since there is a correlation between lag and variability, the variability vs. peak luminosity relation \\citep{rei01} might require similar re-calibration. The simplest explanation for the large number of long-lag GRBs found in the BATSE database is that they are GRBs viewed from off the jet axis. The tightly-beamed, variable, short-lag internal shocks would not be seen from large viewing angles, whereas the more widely-beamed, smooth, long-lag external shocks would be. However, it is difficult to reconcile this explanation at the present time with some of the other similar phenomena associated with GRBs (e.g. x-ray rich GRBs, Gamma-Ray Flashes, and Orphan Afterglows)." }, "0403/astro-ph0403683_arXiv.txt": { "abstract": " ", "introduction": "\\label{intro} We present here a brief review of energy and power spectra, radiative processes giving rise to them, spectral states, and some aspects of variability of black-hole binaries. Our emphasis is on interpretation of the X-ray and soft \\g-ray (hereafter X\\g) observations in terms of simple physical models and unifying various aspects of the wealth of the present observational data. We also discuss millisecond flares recently discovered from Cyg X-1\\cite{gz03}, where we present some important new results. We refer the reader to Ref.~\\citen{mr04} for an exhaustive review of observational aspects of black-hole binaries, and, e.g., to Refs.\\ \\citen{d02,z99,b04} and \\citen{z00} for some other relevant reviews. In our presentation, we use the so-called $\\nu F_\\nu$ representation for both energy and power spectra. The motivation for this choice is provided by a trivial mathematic consideration. Namely, any differential dependence, e.g., flux per unit energy, ${\\rm d}F/{\\rm d}E$, equivalently expressed as $F(E)$ or $F_E$, can be plotted as $F(E)$ vs $E$. However, when we plot $F$ vs.\\ $\\log E$, it should be $F(\\log E)={\\rm d}F/{\\rm d}\\log E$, i.e., $F(E) ({\\rm d}\\,\\log E/{\\rm d}E)^{-1}\\propto E F(E)$. This gives us $F$ per a logarithmic interval of $E$. Graphically, $EF(E){\\rm d}\\log E$ represents the area under the curve, $F$, in a plot vs $\\log E$. Then, a peak in this representation shows us at which photon energy range most of the power is radiated, or which frequency range corresponds to most of variability of a source. An added benefit of this representation is that plots vs.\\ photon energy or wavelength become identical, as $\\lambda F(\\lambda) = E F(E)$. We also plot the energy spectra with the same length per decade on each axis, in order to enable direct comparison of the shape spectra on different figures. On the other hand, it is common in X-ray astronomy to plot photon flux vs.\\ energy. However, this makes it incompatible with most of other branches of astronomy (e.g., those from the radio to the UV), where energy flux is used. Also, energy is a much more fundamental quantity in physics than photon number, with many physical laws concerning the former, not the latter (as photon number is not conserved in most of physical processes). Indeed, all quantities of radiative transfer, e.g., flux, specific and average intensity, radiation pressure, are defined in terms of energy. A practical consequence of plotting photon number per unit energy on a logarithmic plot is that such spectra look usually extremely steep, covering many orders of magnitude on the vertical axis, appear very similar to each other, and make it very difficult to actually see any spectral features. This is due to the artificial steepening of such a spectrum by $E^{-2}$, described above. It is also very common in X\\g\\ astronomy to plot instrumental counts instead of photons or energy. The motivation for it is that high-energy instruments have rather limited energy resolution, which makes the process of deconvolution not unique. However, this purist point of view would actually require plotting only instrumental counts, as plotting together predictions of a model and/or residuals is already model-dependent. In fact, residuals given as, e.g., ratios or contributions to $\\chi^2$, are {\\it identical\\/} for either counts vs.\\ model or a deconvolved spectrum vs.\\ model. However, the latter has the enormous advantages of presenting a physical spectrum, allowing comparison with other spectra, and showing features originating in the source. The count-rate plot, though model-independent, is by its nature instrument-dependent, showing mostly features of the instrumental response, not related to the physics of the source. ", "conclusions": "We have reviewed spectral states of black hole binaries and their respective dominant radiative processes. In the hard state, the dominant process is thermal Comptonization of soft photons, most likely blackbody ones from an outer accretion disk. The thermal character of the distribution of the scattering electrons is probably related to the most likely source geometry, consisting of an inner hot accretion flow\\cite{a95,ny95,z98}, see Fig.\\ \\ref{geo}a. The electrons in the flow are powered by energy transfer from the hot ions, which receive most of the available gravitational energy. The stochastic nature of the energy transfer, probably via Coulomb interactions, results in the electron distribution being nearly thermal. In addition, the cold medium gives rise to reprocessing, including Compton reflection and fluorescence. Relativistic broadening observed in some cases indicates that, in addition to an outer disk, there is some cold medium close to the black hole, e.g, from collapse of the innermost hot flow\\cite{y01}. The variability pattern observed on long timescales indicates that the dominant driver of the variability is changing irradiation of the hot flow by soft seed photons from the outer cold disk\\cite{z02,z03}. \\begin{figure}[t!] \\centerline{\\includegraphics[width=11cm]{hard.ps}} \\vskip 0.7cm \\centerline{\\includegraphics[width=11cm]{soft.ps}} \\caption{(a) A schematic representation of the likely geometry in the hard state, consisting of a hot inner accretion flow surrounded by optically-thick accretion disk. The hot flow consitutes the base of the jet (with the counter-jet omitted from the figure for clarity). The disk is truncated far away from the minimum stable orbit, but it overlaps with the hot flow. The soft photons emitted by the disk are Compton upscattered in the hot flow, and emission from the hot flow is partly Compton-reflected from the disk. (b) The likely geometry in the soft state consisting of flares/active regions above an optically-thick accretion disk extending close to the minimum stable orbit. The soft photons emitted by the disk are Compton upscattered in the flares, and emission from the flares is partly Compton-reflected from the disk\\cite{z02}. } \\label{geo} \\end{figure} In the soft states, Compton scattering is by a hybrid, thermal/nonthermal electron distribution. The scattering plasma forms, most likely, active regions above the surface of an accretion disk, which emit the energetically dominant blackbody emission, see Fig.\\ \\ref{geo}(b). The active regions are powered by energy transfer from inside the disk, most likely via magnetic buoyancy and subsequent magnetic field annihilation, leading then to acceleration of particles. This is inherently a nonthermal process, which explains the observed nonthermal photon spectra, showing no high-energy cutoff up to high energies. The ratio of the power in the active regions to the disk power increases in the sequence of the soft states: ultrasoft, high, intermediate/very high. Thus, the last state is characterized by a powerful corona surrounding most of the disk, in a modification of the picture of Fig.\\ \\ref{geo}(b). In a range of $L/\\ledd$, either the hard/hot and soft/cold accretion flows are possible. This leads to hysteresis in the long-term lightcurve of black-hole binaries, and the corresponding long-term limit cycle. A very interesting new phenomenon in black-hole binaries is that of millisecond flares, extreme events occurring on timescales comparable to $GM/c^3$. In particular, the strongest flare discovered from Cyg X-1 appears to be the most extreme event detected from black-hole accretion yet." }, "0403/astro-ph0403611_arXiv.txt": { "abstract": "Using the deepest and finest resolution images of the Universe acquired with the Hubble Space Telescope and a similar image taken 7 years later for the Great Observatories Origins Deep Survey, we have derived proper motions for the point sources in the Hubble Deep Field--North. Two faint blue objects, HDF2234 and HDF3072, are found to display significant proper motion, 10.0 $\\pm$ 2.5 and 15.5 $\\pm$ 3.8 mas yr$^{-1}$. Photometric distances and tangential velocities for these stars are consistent with disk white dwarfs located at $\\sim$ 500 pc. The faint blue objects analyzed by Ibata et al. (1999) and Mendez \\& Minniti (2000) do not show any significant proper motion; they are not halo white dwarfs and they do not contribute to the Galactic dark matter. These objects are likely to be distant AGN. ", "introduction": "Major observational campaigns have searched for dark matter in the form of massive compact halo objects (MACHOs) using microlensing events (e.g. Alcock et al 1997; Afonso et al. 2003; Udalski et al. 1992). The detection of 13--17 microlensing events toward the Large Magellanic Cloud during 6 years by the MACHO collaboration implies that a significant fraction (20$\\%$) of the halo of the Galaxy may be in the form of compact halo objects (Alcock et al. 2000). The time scale of these lensing events eliminates the possibility of MACHOs having substellar masses. The MACHO collaboration finds a most probable mass of 0.5 \\msun\\, which supports the idea of a massive halo comprised of baryonic matter in the form of low luminosity white dwarfs (Kawaler 1996). Recent observations by the EROS group provide further evidence that less than 25\\% of a standard dark matter halo can be composed of objects with a mass between 2 $\\times$ 10$^{-7}$ \\msun\\ and 1 \\msun\\ (Afonso et al. 2003). Halo white dwarf stars are expected to have large proper motions as a result of their high velocities relative to the Sun. HST proper motion studies of the Globular Cluster NGC 6397 showed that most of the required dark matter in the solar vicinity can be accounted for by a population of old white dwarfs representing the thick disk and halo of the Galaxy (Mendez 2002). Claims by Oppenheimer et al. (2001) and Ibata et al. (2000) that they had found a significant population of halo white dwarfs from kinematic surveys are tantalizing. Their discoveries seemed to be consistent with earlier findings of an old population of white dwarfs in the Hubble Deep Field (Mendez \\& Minniti 2000). However, further analysis by several groups showed that the sample of Oppenheimer et al. (2001) could also be interpreted as the tail of a kinematically warmer white dwarf component, better explained by the thick disk population of the Galaxy (Reid et al. 2001; Reyle et al. 2001; Mendez 2002; Bergeron 2003). The Hubble Deep-Field (HDF) provides a unique window on the Universe (Williams et al. 1996; Flynn et al. 1996). The extreme depth of the HDF provides an unprecedented advantage to find faint stellar objects as well as to study very distant galaxies. The advantage of going deep is that it allows us to search for faint stellar components of the Galaxy in the regions of the color--magnitude diagram that are devoid of any contamination by standard Galactic stars. The lack of ordinary disk stars is due to the finiteness of the Galaxy (Flynn et al. 1996). Mendez \\& Minnitti (2000) claimed that the faint blue objects found in the HDF--North and HDF--South are Galactic stars based on the observed number of blue sources and extragalactic sources in the two fields. Independent proper motion measurements for five of these faint blue sources by Ibata et al. (1999) suggested that they are cool halo white dwarfs which could account for the entire missing mass in the solar neighborhood. Third epoch data on these five objects, however, did not show any significant proper motion (R. Ibata, private communication; Richer 2001). We use the original Hubble Deep Field -- North data and images of the same field taken 7 years later for the Great Observatories Origin Deep Survey (GOODS) to measure proper motions of the point sources analyzed by Ibata et al. (1999) and Mendez \\& Minniti (2000). ", "conclusions": "The nature of the faint blue objects in the Hubble Deep Field may be crucial to understanding the contribution of low luminosity halo white dwarfs to micro-lensing events and the dark matter content of the Galaxy. Apparent proper motions for 5 faint blue objects (Ibata et al. 1999) was enough to explain the entire missing mass in the halo of the Milky Way. Mendez \\& Minniti (2000) claimed that the faint blue objects are white dwarf stars located at heliocentric distances of up to 2 kpc and belong to the Galactic halo. They found a local halo white dwarf mass density of 4.64 $\\times$ 10$^{-3}$ \\msun\\ pc$^{-3}$, which would account for about 30--50$\\%$ of the dark matter in the Galaxy. With the advantage of a 7--year baseline, we are able to place better limits on the proper motion measurements of the faint blue objects. Using the proper motion information, we also derived distances and tangential velocities for these objects. Figure 7 shows the tangential velocities and distances for objects brighter than $V\\approx27$ assuming that they are main sequence stars or DA white dwarfs. All of the main sequence stars exhibit halo kinematics and distances, whereas all of the likely white dwarfs exhibit disk kinematics and distances. Following Gilmore, King, \\& van der Kruit (1989; see also von Hippel \\& Bothun 1990) we use the analytical form of the density profile for the thin disk and thick disk \\begin{equation} \\frac{\\nu_{0}(z)}{\\nu_{0}(0)}= {0.96}\\ {e^{-z/250 pc}} + {0.04}\\ {e^{-z/1000 pc}} \\end{equation} with a local normalization of 0.11 \\msun\\ pc$^{-3}$ (Pham 1997). We use the form \\begin{equation} \\nu_{halo}(r) \\propto \\frac{exp[-7.669\\ (R/R_{e})^{(1/4)}]}{(R/R_{e})^{(7/8)}} \\end{equation} for the halo (Young 1976), where $R$ is the distance from the Galactic center, and R$_{e}$ is the scale factor. $R$ is related to the distance $r$ from the observer to a star by \\begin{equation} R^{2} = R_{0}^{2} + r^{2} - 2 r R_{0}\\ {\\rm cos}b\\ {\\rm cos}l \\end{equation} with $R_{0}$ the solar Galactocentric distance, and {\\it b} and {\\it l} the Galactic coordinates for the HDF--North. We use R$_{0}=$7.8 kpc (Gilmore, King, \\& van der Kruit 1989), R$_{e}=$2.7 kpc (de Vaucouleurs \\& Pence 1978) and a local normalization for the halo of (1/800) $\\times$ 0.11 \\msun\\ pc$^{-3}$ (Chen et al. 2001; Gilmore, King, \\& van der Kruit 1989). Using equations 2, 3, and 4, we calculated the expected number of stars in the HDF. We expect to find 2 thin disk, 3 thick disk, and 11 halo objects in the HDF--North. We have also used Reid \\& Majewski (1993) star count models to predict the number of stars in the HDF--North. We found that 2 thin disk, 4 thick disk, and 14 halo objects are expected in the HDF--North. Both simple analytical models and more sophisticated star count models, when extrapolated to the photometric depth of the HDF, predict similar number of stars (16--20) in the HDF--North. There are 14 stars brighter than $V=27$ and 17 objects fainter than $V=27$ classified as stars by SExtractor. The observed number of stars and the predictions of star count models are in good agreement for $V\\la27$ (see also Mendez et al. 1996 and Mendez \\& Minniti 2000). On the other hand, there seems to be an excess of point sources in the Hubble Deep Field -- North for $V\\ga27$. Unfortunately, SExtractor classification cannot be trusted at these magnitudes. Furthermore, we did not detect significant proper motion for all but two of these objects. The two faint, possibly moving objects, HDF774 and HDF1816, may be halo white dwarfs. One of the problems with any analysis using these objects is that the observations are beyond the completeness limit, and any calculation based on them is subject to a significant completeness correction. The rest of the objects fainter than $V=27$ are probably extragalactic objects (see section 3.2). The five faint blue objects analyzed by Mendez \\& Minniti (2000) do not exhibit any significant proper motion; they are not halo white dwarfs. These objects do not account for the MACHO optical depth and are not the source of the Galactic dark matter. Their stellar nature is not confirmed either. The colors of HDF161 are consistent with our QSO simulations. The faint blue objects may be distant AGN. Holberg et al. (2002) used a local sample of white dwarfs complete out to 13 pc, and found the local mass density of white dwarf stars to be 3.4 $\\pm$0.5 $\\times$ 10$^{-3}$ \\msun\\ pc$^{-3}$. Using this normalization factor in equations 2 and 3, we estimate the expected number of white dwarfs in the Hubble Deep Field. We expect to find 0.05 disk white dwarfs, 0.09 thick disk white dwarfs, and 0.33 halo white dwarfs in the Hubble Deep Field North. We have also used Reid \\& Majewski (1993) star count models to predict the number of white dwarfs in the HDF. The results are roughly consistent: 0.10 disk, 0.25 thick disk, and 0.5 halo white dwarfs are expected. We have discovered two likely white dwarfs, HDF2234 and HDF3072, brighter than $V=27$ in the HDF--North. They are located at distances of $\\sim$ 500 pc and have tangential velocities $\\sim$30 km s$^{-1}$. Their kinematic properties are consistent with being thin disk or thick disk objects (see Table 3 and Figure 7). The expected number of thin disk + thick disk white dwarfs is found to be 0.14 -- 0.35. We have found 6 to 14 times more disk white dwarfs in the HDF--N than expected from the models. Assuming Poisson statistics, the probability of finding two white dwarfs is 1\\% if the expected number of white dwarfs is 0.14 and 4\\% if the expected number of white dwarfs is 0.35. The number of disk + thick disk white dwarfs may be substantially underestimated. Due to small number statistics, however, this statement is only a 2--3 $\\sigma$ result and it heavily depends on the fact that HDF2234 and HDF3072 are white dwarfs. Follow--up spectroscopy of these two objects is needed to confirm this result. Mendez \\& Minniti (2000) have found 22 Galactic stars and 10 faint blue objects in the Hubble Deep Field -- South. A natural test to check the space density of disk and halo white dwarfs would be to obtain second epoch observations of the HDF--South to find high proper motion objects. Also, the HST/ACS Ultra--Deep Field observations of the Chandra Deep Field -- South will be useful to search for faint blue objects at fainter magnitudes and to improve the morphological classification of these objects at brighter magnitudes. The Ultra--Deep Field will be $\\sim$ 1.5 mag deeper than the HDF and HDF--South (Beckwith et al. 2003)." }, "0403/astro-ph0403427_arXiv.txt": { "abstract": "The photon emissivity from the bremsstrahlung process $e^-e^-\\rightarrow e^-e^-\\gamma$ occuring in the electrosphere at the bare surface of a strange quark star is calculated. For surface temperatures $T< 10^{9}$K, the photon flux exceeds that of $e^+e^-$ pairs that are produced via the Schwinger mechanism in the presence of a strong electric field that binds electrons to the surface of the quark star. The average energy of photons emitted from the bremsstrahlung process can be 0.5 MeV or more, which is larger than that in $e^+e^-$ pair annihilation. The observation of this distinctive photon spectrum would constitute an unmistakable signature of a strange quark star and shed light on color superconductivity at stellar densities. \\bigskip \\noindent PACS: 26.60.+c, 95.30.Cq, 97.60.Jd ", "introduction": "\\label{sec_intro} Recently, it has been pointed out that thermal emission from the bare surface of a strange quark star, due to both photons and $e^+e^-$ pair production, can produce luminosities well above the Eddington limit ($\\sim 10^{38}~{\\rm erg~sec}^{-1}$) for extended periods of time, from about a day to decades, depending on the superconducting phase of quark matter~\\cite{Page02}. The spectrum of emitted photons is significantly different from that of a normal cooling neutron star ($30 2000$) (CBI, Pearson et al. 2003; ACBAR, Kuo et al. 2002). The best candidate for this excess (if confirmed) is the SZ effect signal in galaxy clusters. We expect the excess in power due to galaxy clusters to be observed at some point at scales smaller than $\\approx$ few arcmin (or $\\ell > 3000$) although the exact value is model dependent. However, the amplitude of the excess claimed by recent experiments is larger than what it is expected for the current most fashionable models. In fact, if that excess is confirmed to be due to galaxy clusters, it would require high values for the normalization of the matter power spectrum ($\\sigma_8 \\approx 1$) (e.g Bond et al. 2002). These high values of $\\sigma _8$ would contradict the values derived directly from observations of galaxy clusters (e.g Efstathiou et al 2002). Another contribution to the excess in power could be due to unresolved non-subtracted point sources. It is expected to be more important for experiments at low frequencies in the microwave band (like CBI) than at frequencies around 150 GHz (like ACBAR). Since a strong contribution from radio sources is expected in the former case and in the later the extragalactic point source contribution is at the lowest level along the microwave band, a more significant residual after the subtraction/estimation analysis can be present at centimeter rather than millimeter wavelengths. Moreover, the effect of clustering of point sources should be considered in the estimation of the residual (Toffolatti et al. in preparation).\\\\ An interesting aspect of the excess in power is that no compact sources (clusters or point sources) are clearly seen directly in the data, thus suggesting that the origin of this excess may be due to a different nature other than the compact sources. One has only to realize that the power spectrum is an averaged quantity over the surveyed area and that the compact (bright) sources are not distributed over the whole area. As a consequence, many pixels will enter in the average with negligible values. On the other hand, the CMB does distribute over the whole surveyed area. This implies that if the power spectrum of compact sources dominates the power spectrum of the CMB at small scales, then those sources should be seen in the map at small scales. Then, where are they ? The debate is still open and it is not clear whether the excess is due to compact sources or is just a systematic effect.\\\\ In this paper we will study the SZ contribution to the excess found in the power spectrum and its cosmological implications. In particular we will analyze the Gaussian deviations introduced in the wavelet coefficients. Since the CMB has not shown (up to date) any departures from Gaussianity at least at those small scales (for larger scales see Vielva et al. 2003) therefore any non-Gaussian detection would point out the presence of compact sources. The analysis presented in this paper will be useful to confirm or reject the hypothesis that the excess is due to galaxy clusters in current CMB data. The analysis would be also a desirable consistency check with the near future incoming data sets. \\\\ The reader will find very interesting for instance the pioneering work of Aghanim \\& Forni (1999) and the more recent of Rubi\\~{n}o-Martin \\& Sunyaev (2003) where similar questions are considered. ", "conclusions": "We have presented a small (but interesting) excercise showing how current data (ACBAR CMB5 field) is expected to have non-Gaussian signals if the claimed excess in the power spectrum is fully due to the SZ effect. With this excercise we want to send a brief but clear note to the community. Current data should be checked for non-Gaussian signals at small scales before any excess in the power spectrum is claimed as due to the SZ effect. The potentiality of this approach is based on the fact that, up to date, no intrinsic deviations from Gaussianity have been found in the CMB at these small scales. Any deviations would indicate the presence of compact sources which are expected to be the dominant foreground at arcmin scales.\\\\ We have presented the result of the level of non-Gaussianity we should expect in ACBAR-like data if a significant excess in power is due to galaxy clusters. Given the uncertainties in the estimation of the power spectrum by ACBAR and CBI, we explore two models which are in the lower and upper limit of these uncertainties and we find that for the model in the upper limit we should expect a clear non-Gaussian signal at MHW scales around 3 arcmin (the Gaussian and A-model distributions in figures \\ref{fig_1} and \\ref{fig_2} overlap by less than 1 \\%). In the other case (lower limit B-model), we should expect a non-Gaussian signal in $\\approx 75 $\\% of the cases if we look at the kurtosis and in $\\approx 80 $\\% if we look at the skewness (both at $> 99$ \\% confidence limit). By performing this analysis on the ACBAR data (for example), one could confirm or reject several galaxy cluster models. \\\\ Although we have based our simulations on galaxy clusters, a similar argument could be made if the excess is due to point sources. A way to distinguish between both would be the sign of the skewness of the MHW coefficients: a positive sign would point out to point sources whereas a negative one would imply the presence of galaxy clusters (see Rubi\\~{n}o-Martin \\& Sunyaev 2003). For experiments where the level of non-Gaussianity is expected to be small, a more efficient manner to discriminate between the CMB and galaxy clusters is by defining a Fisher discriminant which contains information about both the skewness and kurtosis and also at several scales (Mart\\'\\i nez-Gonz\\'alez et al. 2002). These techniques will prove to be very useful with the near future data. We would like to conclude by being provocative by insisting that actual data should be checked for non-Gaussian signals before suggesting that the excess is due to the SZ effect. If the data is found to be compatible with Gaussianity, then the systematics may be the reason for that excess. For instance, the power spectrum may have been overestimated at small scales if the covariance matrix of the instrumental noise is slightly undetermined at those small scales. Although we are not the first suggesting that non-Gaussianity studies are useful to detect SZ signatures (See Aghanim \\& Forni 1999 or Rubi\\~{n}o-Martin \\& Sunyaev 2003), we are the first in predicting that if a significant excess claimed by recent experiments is due to the SZ effect, in general, non-Gaussian signals should be found with a high significance in at least one of those data sets, the CMB5 field of ACBAR. The reader may consider our approach too simplistic but we think is important to highlight our main point with simple arguments, current data should be checked for non-Gaussian signals." }, "0403/astro-ph0403082_arXiv.txt": { "abstract": "We discuss the propagation of electromagnetic waves on a rectangular lattice of polarizable point dipoles. For wavelengths long compared to the lattice spacing, we obtain the dispersion relation in terms of the lattice spacing and the dipole polarizabilities. We also obtain the dipole polarizabilities required for the lattice to have the same dispersion relation as a continuum medium of given refractive index $m$; our result differs from previous work by Draine \\& Goodman (1993). Our new prescription can be used to assign dipole polarizabilities when the discrete dipole approximation is used to study scattering by finite targets. Results are shown for selected cases. ", "introduction": "The discrete dipole approximation (DDA) provides a flexible and general method to calculate scattering and absorption of light by objects of arbitrary geometry \\citep{Drai88,DrFl94}. The approximation consists of replacing the continuum target of interest by an array of polarizable points, which acquire oscillating electric dipole moments in response to the electric field due to the incident wave plus all of the other dipoles. Broadly speaking, there are two criteria determining the accuracy of the approximation: \\begin{itemize} \\item the interdipole separation $d$ should be small compared to the wavelength of the radiation in the material ($|m|kd < 1$, where $m$ is the complex refractive index, and $k=\\omega/ c$ is the wavenumber {\\it in vacuo}); \\item the interdipole separation $d$ should be small enough to resolve structural dimensions in the target. \\end{itemize} With modern workstations, it is now feasible to carry out DDA calculations on targets containing up to $\\sim10^6$ dipoles \\citep{DrFl94}. In addition to calculation of scattering and absorption cross sections, the DDA has recently been applied to computation of forces and torques on illuminated particles \\citep{DrWe96,DrWe97} If the dipoles are located on a cubic lattice, then in the limit where the interdipole separation $kd\\rightarrow0$, the familiar Clausius-Mossotti relation (see, e.g, \\citet{Ja62}) can be used to determine the choice of dipole polarizabilities required so that the dipole array will approximate a continuum target with dielectric constant $\\epsilon$. \\citet{Drai88} showed how this estimate for the dipole polarizabilities should be modified to include radiative reaction corrections, and \\citet{DrGo93} derived the $O[(kd)^2]$ corrections required so that an infinite cubic lattice would have the same dispersion relation as a continuum of given dielectric constant. Nearly all DDA calculations to date have assumed the dipoles to be located on a cubic lattice. If, instead, a rectangular lattice is used, it will still be possible to apply FFT techniques to the discrete dipole approximation, in essentially the same way as has been done for a cubic lattice \\citep{GoDF91}. However, the ability to use different lattice constants in different directions might be useful in representing certain target geometries. The objective of the present report is to obtain the dispersion relation for propagation of electromagnetic waves on a rectangular lattice of polarizable points. This dispersion relation will be expanded in powers of $(kd)$, where $d$ is the characteristic interdipole separation. For a lattice of specified dipole polarizabilities, this will allow us to determine the dispersion relation for electromagnetic waves propagating on the lattice when $(kd)^2\\ll 1$. Alternatively, if we require the lattice to propagate waves with a particular dispersion relation, inversion of the lattice dispersion relation will provide a prescription for assigning dipole polarizabilities when seeking to approximate a continuum material with a rectangular lattice of polarizable points. ", "conclusions": "The principal results of this paper are as follows: \\begin{enumerate} \\item We have derived the exact mode equation (\\ref{eq:mode_equation}) for propagation of electromagnetic waves through a rectangular lattice of polarizable points. \\item If the electric polarizability tensor $\\alpha$ for the lattice points is specified, then we can solve the mode equation to obtain the dispersion relation for waves propagating on the lattice. Alternatively, if we wish the lattice to have the same dispersion relation as a continuum medium of complex refractive index $m$, then we can solve the mode equation for the appropriate polarizability tensor. \\item In the case of a cubic lattice, our derived ``lattice dispersion relation'' polarizability tensor (\\ref{eq:alpha_cubic}) is diagonal but anisotropic, and differs from the isotropic polarizabilities obtained by DG93. \\item The new polarizability prescription has been tested in DDA scattering calculations for spheres. The resulting cross sections have accuracies comparable to those obtained with the DG93 prescription. \\end{enumerate}" }, "0403/astro-ph0403207_arXiv.txt": { "abstract": "{ We report on a search for low-energy neutrino (antineutrino) bursts in correlation with the 8 time coincident events observed by the gravitational waves detectors EXPLORER and NAUTILUS (GWD) during the year 2001. The search, conducted with the LVD detector (INFN Gran Sasso National Laboratory, Italy), has considered several neutrino reactions, corresponding to different neutrino species, and a wide range of time intervals around the (GWD) observed events. No evidence for statistically significant correlated signals in LVD has been found. Assuming two different origins for neutrino emission, the cooling of a neutron star from a core-collapse supernova or from coalescing neutron stars and the accretion of shocked matter, and taking into account neutrino oscillations, we derive limits to the total energy emitted in neutrinos and to the amount of accreting mass, respectively. ", "introduction": "The analysis of the data collected in coincidence by the gravitational wave bar detectors EXPLORER and NAUTILUS during the year 2001 (Astone et al., 2002) shows an excess (8 events against 2.6 expected from the background) when the two detectors are favorably oriented with respect to the Galactic Disc. Moreover, this result comes from the present day most sensitive experiments for the detection of gravitational wave bursts and a search for neutrino bursts in correlation with the 8 GWD events is, therefore, appropriate. A few astrophysical transient sources are indeed expected to produce associated bursts of neutrinos and gravitational waves. It is well known that most of the energy ($99\\%$) released in the gravitational core collapse of a massive star is carried away by neutrino originated both from the matter accretion in the shock and from the cooling of the proto-neutron star (see for example Burrows et al., 1992). Depending on the collapse dynamics, some fraction of the total energy is emitted in GW ( Thorne 1988, Muller 1997 ), asymmetric supernovae in our Galaxy being the best candidate sources for GW bar detectors. Two coalescing neutron stars would also constitute a source for both neutrinos and gravitational waves. From the point of view of GW emission, it is likely that the merging event would produce powerful gravitational wave bursts, and, even if the physics of the merger is not known, there are estimates that, for binary systems of large mass, coalescence waves are likely to be stronger than the inspiral ones. Some amount of the kinetic energy is converted in thermal energy so that the hot remnant would probably emits thermal neutrinos.\\\\ The search for a neutrino burst associated to the events detected by the GWD EXPLORER and NAUTILUS in 2001 has been performed with the LVD apparatus, operating in the INFN Gran Sasso National Laboratory (Italy) since 1992 with the main purpose of searching for neutrinos from gravitational stellar collapses within the whole Galaxy. The paper is planned as follows: in Sect.2 we briefly describe the LVD detector, and we explain the selection of the data. In Sect.3 we present the results of the analysis: a time interval spanning from 12 $h$ preceding each of the 8 events up to 12 hours later, has been scanned, searching for any excess over the statistical fluctuation of the background. Further, a search for a $\\nu$ signal coincident in time with every event has been performed. We conclude in Sect.4, where we discuss the results of the search, taking into account $\\nu$ oscillations, and considering the two processes for $\\nu$ emission, i.e., cooling and accretion. Since we do not find any neutrino burst candidate associated with the 8 GWD mentioned events, in the following assumed scenarios we derive upper limits: \\begin{itemize} \\item on the neutrino flux, without reference to any particular source, \\item on the total amount of energy emitted in neutrinos, in the cooling case, \\item on the accreting mass, in the accretion case. \\end{itemize} ", "conclusions": "We have conducted a search for low-energy antineutrino (neutrino) bursts with the LVD detector in coincidence with the 8 event excess found by the gravitational waves detectors EXPLORER and NAUTILUS during the year 2001.\\\\ We have found no evidence for any statistically relevant signal in LVD, in three different reaction channels (inverse beta decay, charged current and neutral current interactions with $^{12}C$) corresponding to different neutrino species, over a wide range of time durations, for any of the 8 events. Consequently, we have derived $90 \\%$ fluence upper limits to antineutrino and neutrino emission from an average GW event, at different energies in the range of sensitivity of the LVD detector.\\\\ We have then related the result of the search with two possible simplified models for neutrino emission, i.e., ``cooling'' and ``accretion'', deriving limits, on the one side, to the total energy emitted in neutrinos at the source, and, on the other, to the amount of accreting mass. Assuming a source distance $d=10$ kpc, possible candidates as new-born and colliding neutron stars have been excluded by this analysis. This makes even more challenging and interesting the search for a likely astrophysical source for the reported GWD events.\\\\[1cm] {\\it Acknowledgments}. The authors are grateful to the director and the staff of the National Gran Sasso Laboratory for their constant and valuable support. The authors thank precious comments by Francesco Vissani.\\\\[1cm] {\\bf Appendix A. Limits calculation: inverse $\\beta$ decay interactions.}\\\\[0.2cm] The number of $\\bar{\\nu}_e$ interactions due to inverse beta decay in a detector is given by: \\begin{center} \\begin{equation} N_{ev}^{IBD}=M \\cdot N_{p} \\cdot \\int_{Q}^{\\infty} \\frac{dN_{\\bar{\\nu}_e}}{dE_{\\bar{\\nu}_e}} \\cdot \\sigma(E_{\\bar{\\nu}_e}) \\cdot \\varepsilon_n \\cdot \\varepsilon(E_d,E_{th}) dE_{\\bar{\\nu}_e} \\label{Nevcoolp} \\end{equation} \\end{center} where: $M$ is the detector active mass in ton; $N_p=9.36\\cdot10^{28}$ is the number of free protons per ton; $\\varepsilon_n$ is the neutron detection efficiency; $\\varepsilon(E_d,E_{th})$ is the $e^+$ detection efficiency; $E_d=E_{\\bar{\\nu}_e}-Q+2m_ec^2$, with $Q=M_n+m_e-M_p=1.8$ MeV is the positron detectable energy; $E_{th}$ is the detector energy threshold; $\\sigma(E_{\\bar{\\nu}_e})$ is the cross section (Vogel and Beacom, 1999; Strumia and Vissani, 2003); $\\frac{dN_{\\bar\\nu_e}}{dE_{\\bar\\nu_e}}$ is the antineutrino energy spectrum.\\\\ In the case of cooling process it is: \\begin{center} \\begin{equation} \\frac{dN_{\\bar\\nu_e}}{dE_{\\bar\\nu_e}}= \\frac{E_{B}}{4\\pi d^2} \\cdot \\frac{120}{7\\pi^4} \\cdot F_{\\bar\\nu_e} \\end{equation} \\end{center} where $d=10 Kpc$ is the assumed source distance;\\\\ $E_{B} = \\sum_{i}f_{\\nu_i}E_{B}$ is the total energy emitted in neutrinos;\\\\ $F_{\\nu}$ is the term accounting for different $\\nu$ oscillation scenarios\\footnote{NC data are not affected by oscillations. However, the limits from IBD data stay almost the same even assuming that MSW oscillations are completely absent. In fact, on accounting for vacuum oscillations we get $P_{\\bar{e}\\bar{e}}=1-sin^2 2\\theta_{12}/2 \\sim 0.6$ in all scenarios; this is practically the same value implied by MSW oscillations in the scenarios $NH$ and $IH_{non\\ ad.}$, $P_{\\bar{e}\\bar{e}}=cos^2\\theta_{12}\\sim 0.7$.}: \\begin{itemize} \\item for $NH_{ad}$; $NH_{non-ad}$; $IH_{non-ad}$: \\begin{center} \\begin{equation} F_{\\bar{\\nu}_e}= \\frac{f_{\\nu_e}}{T_{\\bar{\\nu}_e}^4}|U_{e1}|^2\\frac{E^2}{1+e^{E/ T_{\\bar{\\nu}_e}}} + \\frac{f_{\\nu_x}}{T_{\\bar{\\nu}_x}^4}|U_{e2}|^2\\frac{E^2}{1+e^{E/ T_{\\bar{\\nu}_x}}} \\label{osc1ae} \\end{equation} \\end{center} \\item for $IH_{ad}$: \\begin{center} \\begin{equation} F_{\\bar{\\nu}_e}= \\frac{f_{\\nu_e}}{T_{\\bar{\\nu}_e}^4}|U_{e3}|^2\\frac{E^2}{1+e^{E/ T_{\\bar{\\nu}_e}}} + \\frac{f_{\\nu_x}}{T_{\\bar{\\nu}_x}^4}(1-|U_{e3}|^2)\\frac{E^2}{1+e^{E/ T_{\\bar{\\nu}_x}}} \\label{osc2ae} \\end{equation} \\end{center} \\end{itemize} with: $|U_{e1}|^2 \\approx cos^2\\theta_{12}=0.67$, $|U_{e2}|^2 \\approx sin^2\\theta_{12}=0.33$, $|U_{e3}|^2 \\geq 10^{-4}$ for the adiabatic case and $|Ue3|^2 \\leq 10^{-6}$ for the non adiabatic one (M. Apollonio et al., 1999) (J.Bahcall and C.Pena-Garay, 2003).\\\\ In the case of mass accretion process it is (Loredo and Lamb, 2002): \\begin{center} \\begin{equation} \\frac{dN_{\\bar\\nu_e}}{dE_{\\bar\\nu_e}}= \\frac{1}{4\\pi d^2} \\cdot A_a \\cdot Y_n \\cdot M_{hot} \\cdot F'_{\\bar\\nu_e} \\end{equation} \\end{center} where:\\\\ $M_{hot}$ is the mass of hot emitting material;\\\\ $Y_n$ is the neutron fraction;\\\\ $A_a=\\frac{1+3g^2_A}{8}\\frac{\\sigma_{0}c}{m_n(m_ec^2)^2}\\frac{8\\pi}{(hc)^3}$, with $g_A=1.254$, $\\sigma_{0}=1.7 \\cdot 10^{44} cm^2$;\\\\ and with respect to $F'_{\\nu}$: \\begin{itemize} \\item for $NH_{ad}$; $NH_{non-ad}$; $IH_{non-ad}$: \\begin{equation} F'_{\\bar{\\nu}_e}= |U_{e1}|^2\\frac{E^4}{1+e^{E/ T_{\\bar{\\nu}_e}}} \\label{osc3ae} \\end{equation} \\item for $IH_{ad}$: \\begin{equation} F'_{\\bar{\\nu}_e}=|U_{e3}|^2\\frac{E^4}{1+e^{E/ T_{\\bar{\\nu}_e}}} \\simeq 0 \\label{osc4ae} \\end{equation} \\end{itemize} \\vspace{0.2cm} {\\bf Appendix B. Limits calculation: neutral current interactions}\\\\[0.2cm] The number of interactions in the detector due to the neutral current is given by: \\begin{center} \\begin{equation} N_{ev}^{NC}=M \\cdot N_{C} \\cdot \\varepsilon_C \\cdot \\int_{15.1\\ {\\rm{MeV}}}^{\\infty}[\\frac{dN_{\\bar{\\nu}_i}}{dE_{\\bar{\\nu}_i}} \\sigma(E_{\\bar{\\nu}_i}) + \\frac{dN_{\\nu_i}}{dE_{\\nu_i}} \\sigma(E_{\\nu_i})] dE \\label{Nevcoolcn} \\end{equation} \\end{center} where: $N_{C}=4.24\\cdot10^{28}$ is the number of $^{12}C$ nuclei per ton; $\\varepsilon_C$ is the detector efficiency for $15.1$ MeV gamma; $\\sigma (E_{\\nu})$ is the cross section (M. Fukugita et al., 1988). \\\\ The neutrino energy spectrum in the case of cooling process is: \\begin{center} \\begin{equation} \\frac{dN_{\\bar\\nu_i}}{dE_{\\bar\\nu_i}}=\\frac{dN_{\\nu_i}}{dE_{\\nu_i}}= \\frac{E_{B}}{4\\pi d^2} \\cdot \\frac{120}{7\\pi^4} \\cdot F_{i} \\end{equation} \\end{center} with $F_i= \\frac{f_{\\nu_e}}{T_{{\\nu}_e}^4}\\frac{E^2}{1+e^{E/ T_{{\\nu}_e}}} + 2 \\cdot \\frac{f_{\\nu_x}}{T_{{\\nu}_x}^4}\\frac{E^2}{1+e^{E/ T_{{\\nu}_x}}}$\\\\ while, for the mass accretion case, we considered all the events as $\\bar\\nu_e$ and we used: \\begin{center} \\begin{equation} \\frac{dN_{\\bar\\nu_i}}{dE_{\\bar\\nu_i}}=\\frac{dN_{\\bar\\nu_e}}{dE_{\\bar\\nu_e}}= \\frac{1}{4\\pi d^2} \\cdot A_a \\cdot Yn \\cdot M_{hot} \\cdot \\frac{E^4}{1+e^{E/ T_{\\bar \\nu_e}}} \\end{equation} \\end{center}" }, "0403/astro-ph0403031_arXiv.txt": { "abstract": "In this work we investigate the evolution of matter density perturbations for two different quintessence models. One of them is based on the Einstein theory of gravity, while the other is based on the Brans-Dicke scalar tensor theory. We show that it is possible to constrain the parameter space of the models using the determinations for the growth rate of perturbations derived from data of the 2-degree Field Galaxy Redshift Survey. ", "introduction": "In the past few years, it has become apparent that the energy budget of our universe is dominated by an unknown component called \"dark energy\". The WMAP table of \"best\" cosmological parameters \\cite{lambda}, for instance, gives a $0.73\\pm0.04$ abundance for it, and a value of its equation of state $\\omega<-0.78$. In order to relieve some problems of the popular $\\Lambda CDM$ model (like the fine tuning issue), a dynamical $\\Lambda$-term has been proposed as representative of the dark energy. It's more popular version is a slowly rolling scalar field named quintessence. Many alternative cosmological models have been proposed, and indeed it is a challenge the work of ruling out all the \"incorrect ones\" on observational grounds. For instance many different potentials for these self interacting scalar fields (quintessence) have been proposed. However, it is obvious the importance of the observational exploration of the proposed cosmological models, and in this paper we give a further step in this direction. In \\cite{Isra1} two cosmological models are proposed. In both cases two fluids fill the universe: a background fluid of ordinary matter and a self-interacting scalar-field fluid accounting for the dark energy, both in Einstein's theory and in Brans-Dicke gravity. A linear relationship between the Hubble expansion parameter and the time derivative of the scalar field is used to derive exact cosmological attractor-like solutions. And a priori assumptions about the functional form of the potential or the scale factor behavior are not necessary. All these features render two very interesting models that motivated us to proceed with the observational check of them. In the original paper, the authors made a first check with some observational facts, related with Cosmic Microwave Background (CMB), nucleosynthesis, structure formation (restriction on quintessence density during galaxy formation epoch) and type Ia supernovae (SN Ia). Now we proceed the observational checking considering another aspect of structure formation: the galaxy motions and clustering, i. e., the evolution of density perturbations in the Universe. In the past few years, observations of the large scale structure of the Universe have improved greatly. The development of fiber-fed spectrographs that can simultaneously measure spectra of hundreds of galaxies has provided large redshift surveys such as the 2-degree Field Galaxy Redshift Survey (2dFGRS) and the Sloan Digital Sky Survey (SDSS). In particular, the Anglo-Autralian Telescope of the 2dFGRS has obtained the redshift of a quarter million galaxies. This collaboration has produced abundant data and technical papers \\cite{2dFGRS} about galaxy motions and clustering, and we will refer to some of this, in particular their velocity/density comparisons. The paper is organised as follows: in section II we outline the main characteristics of the models, in section III the main aspects of velocity/density comparisons are exposed and the equation for the growth of perturbations is solved, in section IV the observational check is presented and interpreted, while in section V conclusions are drawn. Finally, references are supplied. ", "conclusions": "In the models presented here we found that the study of perturbation growth is a good tool to constraint the parameter space. Research on the origins and evolution of the large-scale of the universe is one of the hottest topics in cosmology. In this work, we have used the relation between the peculiar velocity field of the galaxies, the growth rate of perturbations and the density bias in galaxy formation to make another step in the observational check of two quintessence models, resulting in a further constrain on the parameter space of the models. We plan in the oncoming future proceed this works using another cosmological probes, like CMB, for instance." }, "0403/astro-ph0403341_arXiv.txt": { "abstract": "Using the method of multiple scales, one can derive an analytic solution that describes the behaviour of weakly coupled, non-linear oscillations in nearly Keplerian discs around neutron stars or black holes close to the 3:2 orbital epicyclic resonance. The solution obtained agrees with the previous numerical simulation. Such result may be relevant to the kilohertz quasi-periodic variability in X-ray fluxes observed in many Galactic black hole and neutron star sources. With a particular choice of tunable parameters, the solution fits accurately the observational data for Sco X-1. ", "introduction": "Many Galactic black hole and neutron star sources in low mass X-ray binaries show both chaotic and quasi-periodic variability in their observed X-ray fluxes. Some of the quasi-periodic oscillations (QPOs) are in the kHz range and often come in pairs $(\\nu_{\\rm upp}, \\nu_{\\rm down})$ of {\\it twin peaks} in the Fourier power spectra (e.g., van der Klis, 2000). In all four black hole sources with twin peak kHz QPOs pairs, $\\nu_{\\rm upp}/\\nu_{\\rm down} = 3/2\\,$ (McClintock \\& Remillard, 2003). For neutron stars sources the ratio of twin-peak frequencies is mostly close to $3/2$ too (Abramowicz, Bulik, Bursa, Klu\\'zniak, 2003). Based on these and other properties of QPOs Klu\\'zniak \\& Abramowicz (2000) concluded that twin peak kHz QPOs are due to a resonance in the accretion disc oscillation modes and they noticed that the specific $3:2$ ratio could be a consequence of strong gravity. According to the standard Shakura-Sunyaev accretion disc model, matter spirals down into the central black hole along stream lines that are located almost on the equatorial plane $\\theta = \\theta_0 = \\pi/2$, and that locally differ only slightly from a family of concentric circles $r = r_0 = const$. The small deviations, $\\delta r = r - r_0$, $\\delta \\theta = \\theta - \\theta_0$ are governed, with accuracy to linear terms, by \\begin{equation} \\label{Equation1} \\delta \\ddot r + \\omega_r^2 \\,\\delta r = \\delta a_r , ~~~~ \\delta \\ddot \\theta + \\omega_{\\theta}^2\\,\\delta \\theta = \\delta a_{\\theta} . \\end{equation} \\noindent Here the dot denotes the time derivative. For purely Keplerian (free) motion $\\delta a_r = 0$, $\\delta a_{\\theta} = 0$ and the above equations describe two uncoupled harmonic oscillators with the eigenfrequencies $\\omega_{\\theta}$, $\\omega_r$. \\noindent We shall start with an argument which appeals to physical intuition and which shows that the resonance now discussed is a very natural, indeed necessary, consequence of strong gravity. Assume that in thin discs, random fluctuations have $\\delta r \\gg \\delta \\theta.$\\footnote{In actual calculations this additional assumption is not made.} Thus, $\\delta r \\delta \\theta$ is a first order term in $\\delta \\theta$ and should be included in the first order equation for vertical oscillations (\\ref{Equation1}). The equation now takes the form, \\begin{equation} \\label{Equation5} \\delta \\ddot \\theta + \\omega_{\\theta}^2\\left [ 1 + h\\,\\delta r \\right ] \\delta \\theta = \\delta a_{\\theta}, \\end{equation} \\noindent where $h$ is a known constant. The first order equation for $\\delta r$ has the solution $\\delta r = A_0 \\cos (\\omega_r \\,t)$. Inserting this in (\\ref{Equation5}) together with $\\delta a_{\\theta} = 0$, one arrives to the Mathieu equation ($A_0$ is absorbed in $h$), \\begin{equation} \\delta \\ddot \\theta + \\omega_{\\theta}^2\\left [ 1 + h \\,\\cos (\\omega_r \\,t)\\right ] \\delta \\theta = 0, \\end{equation} \\noindent that describes the {\\it parametric resonance}. From the theory of the Mathieu equation one knows that when \\begin{equation} {\\omega_r \\over \\omega_{\\theta}} = {\\nu_r \\over \\nu_{\\theta}} = {2 \\over n}, ~~~~n =1,\\,2, \\,3 ..., \\end{equation} \\noindent the parametric resonance is excited (Landau \\& Lifshitz, 1976). The resonance is strongests for the smallest possible value of $n$. Because near black holes and neutron stars $\\nu_r < \\nu_{\\theta}$ (see Figure \\ref{fig1}), the smallest possible value for resonance is $n = 3$, which means that $2\\,\\nu_{\\theta} = 3\\,\\nu_r$. This is an example of a situation in which parametric resonance works in a thin accretion disc and it intuitively explains the observed 3:2 ratio (Klu\\'zniak and Abramowicz, 2002), because, obviously, \\begin{equation} \\nu_{\\rm upp} = \\nu_{\\theta},~~~~\\nu_{\\rm down} = \\nu_r. \\end{equation} \\noindent Of course, in real discs neither $\\delta r = A_0 \\cos (\\omega_r \\,t)$, nor $\\delta a_{\\theta} = 0$ exactly; moreover in realistic situations, for purely geodesic motion ($\\delta a_{\\theta}=\\delta a_r=0$) the system does not show increasing amplitudes (the higher terms prevent this from occurring).\\\\ However, one may expect that because these equations are approximately obeyed in thin discs, the parametric resonance will indeed be excited. Such a resonance was found in numerical simulations of oscillations in a nearly Keplerian accretion disc by Abramowicz et al. (2003). \\begin{figure}[!tb] \\begin{center} \\FigureFile(0.48\\textwidth,0.48\\textwidth){freq.eps} \\end{center} \\caption{Epicyclic frequencies in Schwarzschild's metric: meridional (solid line) and radial one (dashed line). The radius is in units of $r_G$. The vertical lines indicates the radii ($r_{3:2}$,$r_{5:3}$,$r_{4:2}$) where the ratio between the two frequencies is $3:2$ , $5:3$ and $4:2$ (from right to left).} \\label{fig1} \\end{figure} ", "conclusions": "The Taylor expansion of the relativistic geodesics to the third order leads to two coupled harmonic oscillators: the purely geodesics motion is stable, while when it is perturbed there are cases for which parametric resonance may occur.\\\\ Using the method of multiple scales one can highlight that, owing to the curvature (through the way it determines the effective potential), the first allowed resonance between the radial an the vertical epicyclic frequencies is the $3:2$, in agreement with the numerical analysis.\\\\ Moreover the deviation of the slope from $3:2$ is easily explained as a property of such a non-linear resonance\\\\ The analysis of this toy-model reinforces the theory that indeed the observed pairs of QPOs may be due to parametric resonance, and finally to the strong gravitational field alone. \\medskip The work reported here is a part of my 2003 laurea thesis {\\it Risonanza parametrica di dinamiche quasi geodetiche con campi gravitationali forti} at the Physics Department, University of Trieste. I thank my external thesis supervisor ({\\it correlatore}), Marek Abramowicz, for suggesting the subject to me, the many discussions and his constant support . I also thank my local supervisor ({\\it relatore}), Fabio Benatti, for his advice and help and Wlodek Klu\\'zniak for the help in revising this paper. I was working on the problem at the University of Trieste and several other institutions: at SISSA (Trieste, Italy), Chalmers University (G\\\"oteborg, Sweden), UKAFF (Leicester University, England), and the Cargese 2003 Spring School on Black Holes (Cargese, France). I thank all these institutions for hospitality. This work have been supported by the EU grant in Leicester,by MA's VR Swedish goverment grant and by the Polish KBN Grant 2P03D01424. \\appendix" }, "0403/astro-ph0403388_arXiv.txt": { "abstract": "We explore theoretical models of the ionization ratios of the Li-like absorbers \\ion N5, \\ion O6, and \\ion C4, in the Galactic halo. These ions are believed to form in nonequilibrium processes such as shocks, evaporative interfaces, or rapidly cooling gas, all of which trace the dynamics of the interstellar medium. As a useful new diagnostic, we focus on velocity-resolved signatures of several common physical structures: (1) a cooling Galactic fountain flow that rises, cools, and recombines as it returns to the disk; (2) shocks moving toward the observer; (3) a conductive interface with the observer located in the hotter gas. This last geometry occurs with the solar system inside a hot bubble, or when one looks out through the fragmenting top shell of our local bubble blown into the halo as part of the Galactic fountain. In Paper~II, these models are compared to ionization-ratio data from FUSE and {\\it Hubble Space Telescope}. ", "introduction": "The nature and dynamics of the interstellar medium (ISM) of galaxies determines how the energy and matter released by stars are redistributed through the universe. It is thus critical to understand the ISM, and in particular that of the Milky Way, which can be observed with greater sensitivity and resolution than other galaxies. The ISM in the disk of our galaxy consists of several phases, from dense and cold molecular gas to hot and fully ionized. Stars heat and disperse the dense clouds in which they form, and that hot gas cools and recombines eventually to complete the cycle and form more stars. It is particularly interesting to consider the interface between this multiphase medium and intergalactic space, where hot gas is released from supernovae to several kiloparsecs in altitude, forming a hot diffuse medium first proposed by \\cite{spitzer56} as the Galactic corona. The Galactic corona or halo is almost certainly a dynamic object. It is difficult to construct static halo models, because models supported by thermal pressure are thermally unstable. If conduction is sufficiently important to stabilize small-scale instabilities, then the entire (nearly isothermal) halo is unstable to collapse or expulsion as a wind \\citep[see][for stability arguments]{bregman80,field65}. Cosmic-ray supported static halos have been proposed \\citep[e.g.,][]{boulares90}, but there are considerable uncertainties as to how cosmic rays are confined by the Galaxy (the Galactic magnetic field topology in particular), and these models do not explain the high and intermediate velocity clouds. Considering these things, \\citet{shapfield76} first proposed a ``Galactic fountain'' of supernova-heated gas rising buoyantly above the disk until it cools and falls back to the disk. Smooth Galactic fountain models have been constructed by several authors. In particular, \\citet{bregman80} discussed the issues of radial flow in the Galactic gravitational field for a supersonic hot flow which reaches several kiloparsecs height. More recent models \\citep{houck90,breit93} consist of transsonic flows, which require cooler initial temperatures and only rise to 1--2~kpc. These models can help to explain the population of intermediate velocity halo clouds, believed to exist at those lower heights and velocities compared to the more distant high velocity clouds \\citep[e.g.,][]{hivcdist}. The cooling layer certainly fragments into small clouds by Rayleigh-Taylor instability even if no other inhomogeneities exist \\citep[e.g.,][]{berry98}. Superbubbles, worms, and shell-like structures have been observed in \\citep[e.g.,][]{heiles84} the upper disk/low halo (few hundred parsecs altitude), but it is difficult to ascertain whether the hot gas is being expelled into the halo. A related issue is the filling factor of hot gas in the disk, which determines the rate at which hot gas can escape into the halo. Estimates for that filling factor and for the rate of reheating by halo supernovae are based on the evolution of supernova remnants in smooth media, but our current understanding of the ISM is increasingly inhomogeneous and dynamic, so those arguments are probably of limited value \\citep{kahn98}. Recently, the combination of good observations, sophisticated numerical models, and sufficient comprehension of the Galactic fountain may be beginning to allow identification of fountain-like rising structures \\citep{avillez01}. The dynamics of interstellar gas in general and specifically in the Galactic halo may be best understood by studying the hot phase of the ISM (coronal gas at millions of degrees), and gas at temperatures $\\sim$10\\ts{5}~K intermediate to the hot phase and cooler phases (10\\ts{4}~K). Of the different interstellar gas phases, the coronal gas is most directly linked to the main sources of energy in the ISM, supernovae and stellar winds. Slightly cooler gas is most closely linked to transient and dynamical processes. This gas is typically short-lived because the cooling time is short at at 10\\ts{5}~K. The lithium-like ions of common metals, in particular \\ion O6, \\ion N5, and \\ion C4, are sensitive tracers of interstellar gas at several times 10\\ts{5}~K. The resonance absorption lines of these ions are observable with ultraviolet spectrographs on the {\\it Far Ultraviolet Spectroscopic Explorer} (FUSE), and the {\\it Space Telescope Imaging Spectrograph} (STIS) and {\\it Goddard High Resolution Spectrograph} (GHRS) aboard the {\\it Hubble Space Telescope} (HST). Section \\ref{hotmodels} describes previous models of high-ion column densities. Section \\ref{interpretation} describes new models of the dynamical signatures of the Li-like ions and their interpretation. We summarize in \\S~4. ", "conclusions": "We have considered the diagnostic power of velocity-resolved column density ratios in understanding the Galactic halo. Column density ratios of Li-like ions in the Galaxy are useful to diagnose the physical formation mechanism of the gas and to study the interstellar gas cycle, and a survey of these ions can reveal general trends. In Paper II, we present a survey of sightlines observed with FUSE and HST, in which the distribution of N(\\N) and N(\\O) in the halo does not appear to favor a dominant physical production mechanism. Here, we have presented models of interfaces and cooling nonequilibrium gas, focusing on the velocity-resolved N(\\N)/N(\\O) signatures. In particular, we consider conductive interfaces, radiative shocks and supernova remnants, and cooling gas in a Galactic fountain. Typical ion column density ratios are summarized in Table~\\ref{summary}, along with a typical observed slope observed in FUSE and HST data. One important type of hot/cool gas interface for which the ionization-velocity signature has not been discussed is the turbulent mixing layer \\citep{mixing}. The structure of these Li-like ion producing structures is uncertain. It is likely that there is a turbulent cascade of eddies which would wash out any velocity-ionization signature, but there may possibly be large Kelvin-Helmholtz rolls with some coherent velocity signature. \\begin{deluxetable}{lr} \\tabletypesize{\\footnotesize} \\tablecaption{\\label{summary} Summary of models} \\tablecomments{Models predict a slope in N(\\N)/O(\\O) with line-of-sight velocity for observers looking through the structure. Galactic fountain model includes \\citet{bbc} cooling } \\tablehead{ \\colhead{Model} & \\colhead{Slope of Ratio log[N(\\N)/N(\\O)]} } \\startdata \\parbox{3in}{Radiative shock} & $\\sim$-0.0015 dex~(km~s\\ts{-1})\\ts{-1} \\\\ \\parbox{3in}{Mature SNR shell \\citep{shelton98}} & $\\gtrsim$-0.02 dex~(km~s\\ts{-1})\\ts{-1} \\\\ \\parbox{3in}{\\citet{houck90} fountain} & $\\sim$+0.002 dex~(km~s\\ts{-1})\\ts{-1} \\\\ \\tableline Indicative observed slope & -0.0032+/-0.0022(r)+/-0.0014(s) dex~(km~s\\ts{-1})\\ts{-1} \\enddata \\end{deluxetable} The observable velocity-ionization trends are weak, because even very strong trends are washed out by the large thermal width of the gas at different parts of the flow. These trends could be further complicated by thermal instabilities, which likely occur in cooling gas such as in a fountain flow, and can also affect shocks and supernova shells. Fragementation of the gas into parcels with complex density and velocity structure, as seen in the time-dependent shock model of \\citet{sutherland} could complicate the signatures described here. Additional confusion can result when long sightlines pass through multiple structures. As we show in Paper II, the dispersion of N(\\N)/N(\\O), both integrated and velocity-resolved, indicates that no single production scenario known to date can completely explain the Galactic halo. To truly understand the physical production of Li-like ions in the halo, one needs to analyze gas in localized areas of physical space, rather than velocity space. Absorption spectroscopy towards many halo stars with close angular separation and different distances could help to isolate gas at a specific altitude. Similarly, the gas above known superbubble shells or chimneys could be isolated. These observations have a greater chance of distinguishing between models of hot gas production than observations along long lines of sight." }, "0403/astro-ph0403177_arXiv.txt": { "abstract": "{ A sample of 37 Galactic, 10 LMC and 6 SMC cepheids is compiled for which individual metallicity estimates exist and $BVIK$ photometry in almost all cases. The Galactic cepheids all have an individual distance estimate available. For the MC objects different sources of photometry are combined to obtain improved periods and mean magnitudes. A multi-parameter Period-Luminosity relation is fitted to the data which also solves for the distance to the LMC and SMC. When all three galaxies are considered, without metallicity effect, a significant quadratic term in $\\log P$ is found, as previously observed and also predicted in some theoretical calculations. For the present sample it is empirically determined that for $\\log P < 1.65$ linear $PL$-relations may be adopted, but this restricts the sample to only 4 LMC and 1 SMC cepheid. Considering the Galactic sample a metallicity effect is found in the zero point in the $VIWK$ $PL$-relation ($-0.6 \\pm 0.4$ or $-0.8 \\pm 0.3$ mag/dex depending on the in- or exclusion of one object), in the sense that metal-rich cepheids are brighter. The small significance is mostly due to the fact that the Galactic sample spans a narrow metallicity range. The error is to a significant part due to the error in the metallicity determinations and not to the error in the fit. Including the 5 MC cepheids broadens the observed metallicity range and a metallity effect of about $-0.27 \\pm 0.08$ mag/dex in the zero point is found in $VIWK$, in agreement with some previous empirical estimates, but now derived using direct metallicity determinations for the cepheids themselves. ", "introduction": "The importance of the cepheid Period-Luminosity relation ($PL$-relation) has long been recognised and is the basis of the determination of the Hubble constant by Mould et al. (2000) and Freedman et al. (2001). The most important uncertainties in this derivation are the zero point of the $PL$-relation based on an adopted distance modulus (DM) to the LMC, and the adopted metallicity correction. Mould et al. (2000) and Freedman et al. (2001) adopt corrections in the Wesenheit-index ($W$ = $V - 2.55 (V-I)$) of $-0.24 \\pm 0.16$ and $-0.2 \\pm 0.2$ mag/dex, respectively. This metallicity effect is important as the galaxies surveyed by the HST Key Project span a factor of 30 in oxygen abundance (Ferrarese et al. 2000). Theoretical pulsation models lead to different results: linear models (e.g. Sandage et al. 1999, Alibert et al. 1999, Baraffe \\& Alibert 2001) predict a moderate dependence, while non-linear convective models (Bono et al. 1999, Caputo et al. 2000) predict a larger dependence, and in the sense that metal-rich cepheids are fainter (see Table~7 in Groenewegen \\& Oudmaijer 2000, for values at typical periods). Recently, Fiorentino et al. (2002) suggested that there is also a dependence on the Helium abundance. On the observational side the results seem to indicate consistently that metal-rich cepheids are brighter, and various estimates have been given in the literature, $-0.88 \\pm 0.16$ mag/dex ($BRI$ bands, Gould 1994;), $-0.44^{+0.1}_{-0.2}$ mag/dex ($VR$ bands, Sasselov et al. 1997), $-0.24 \\pm 0.16$ mag/dex ($VI$ bands, Kochanek 1997), $-0.14 \\pm 0.14$ mag/dex ($VI$ bands, Kennicutt et al. 1998), $-0.25 \\pm 0.05$ mag/dex ($VI$ bands, Kennicutt et al. 2003), and $-0.21 \\pm 0.19$ in $V$, $-0.29 \\pm 0.19$ in $W$,$-0.23 \\pm 0.19$ in $I$,$-0.21 \\pm 0.19$ mag/dex in $K$ (Storm et al. 2004). The potential drawback or caveat is that no individual abundance determinations of individual cepheids are being used in these studies but rather abundances of nearby H{\\sc ii} regions, or even a mean abundance of the entire galaxy. The present paper aims at investigating the metallicity dependence from the observational side, but using cepheid individual metallicity determinations. This has become possible because of advances in abundance determinations for the Galaxy (Fry \\& Carney 1997, Andrievsky et al. 2002a,b,c, Luck et al. 2003) and LMC (Luck et al. 1998), as well as recent advances in individual distance estimates for Galactic cepheids based on surface-brightness relations (Fouqu\\'e et al. 2003), a Bayesian statistical analysis to solve the surface-brightness equations (Barnes et al. 2003), direct measurement of distances based on combining radial velocity data with interferometric observations (Kervella et al. 2003), and distance determinations of cepheids in open clusters (Tammann et al. 2003). The paper is organised as follows. In Sect.~2 the datasets on individual distance and metallicity determinations for Galactic cepheids are presented and compared. The Magellanic Cloud sample is also described. In Sect.~3 the model and the results are presented. The conclusions and future prospects are outlined in Sect.~4. ", "conclusions": "An attempt has been made, using the currently available data of cepheids with direct metallicity determinations, to quantify the metallicity dependence of the cepheid $PL$-relation. For a purely Galactic sample, the range in metallicity covered is too small to draw any firm conclusions. A metallicity effect of $-0.6 \\pm 0.4$ or $-0.8 \\pm 0.3$ mag/dex in $VWK$ is derived, depending on whether the only Galactic star with a significantly sub-solar metallicity is excluded or included. For the combined sample of Galactic, SMC and LMC cepheids the problem is that most of the MC cepheids have such long periods, in the regime where the $PL$-relation is no longer linear. Restricting the sample in periods to $\\log P < 1.65$ a metallicity effect of about $-0.27 \\pm 0.08$ mag/dex in the zero point is found (in $VIWK$). However, the sample of MC cepheids is presently too small to additionaly solve for a metallicity dependence of the slope in $\\log P$. Formally, the result for the Galactic sample is in agreement with that comprising all three galaxies. This result based on the sample including the MC cepheids is in agreement with other recent empirical estimates (most recently by Kennicutt et al. 2003 and Storm et al. 2004). However, it is stressed that this study is the first to use {\\em direct} cepheid metallicity determinations to arrive at this result. In agreement with Storm et al. (2004) we find that in the range $V$ to $K$ the metallicity effect seems not to depend on wavelength, contrary to theoretical calculations (e.g. Bono et al. 1999), which predicted a decreasing effect towards longer wavelengths. The method here developed (Eq.~(1)) is general and can easily be extended to other functional forms, or to include other galaxies. In principle, the distance to a galaxy and the metallicity dependence can be derived independently. For the present restrictive sample this is not the case however, since the distribution over metallicity is correlated with the parent galaxy. In other words, $\\alpha_2$, $\\Delta_{\\rm LMC}$ and $\\Delta_{\\rm SMC}$ are strongly correlated: a change in the DM to the MCs can be 'compensated for' by a change in the metallicity effect. Therefore, it was decided to not solve for the distance moduli. However, in the second block of results in Table~\\ref{TAB-RES1} (no metallicity effect, but allowing for a quadratic term in $\\log P$) there is some indication that the best fitting DM might be slightly longer than 18.50 for the LMC and slightly shorter than 18.90 for the SMC (although there is some scatter depending on the wavelength). Simply adopting values of 18.55 and 18.80, respectively, within the error bars of the current best independent estimates (Feast 2003, Walker 2003), would result in a value for $\\alpha_2$ of $-0.38 \\pm 0.08$ mag/dex in $V$, that is, it would make the metallicity effect slightly stronger (and similarly for the other bands). It is shown that the error in the independent variable metallicity is a significant contributor to the final error in the metallicity effect as it is larger than the error in the fitting in the dependent variable, but of little effect in the final error in slope and zero point. The mean error in metallicity of the stars in the Galactic plus MC sample is about 0.08 dex. By repeating the Monte Carlo simulations it was derived that for a uniform error bar of 0.05 dex the two sources of error become comparable, and that for a uniform error bar of 0.03 dex or less the error in the metallicity becomes insignificant compared to the fit error. Although the present results seem to indicate a significant metallicity effect, more data are needed to quantify this effect better. An important issue is an internally consistent metallicity scale. The simulations even suggest an accuracy to a level of 0.05 dex or better. However, this seems very difficult to achieve in practice as the error in the metallicity not only depends on the S/N in the spectra and the number of lines, but also on the uncertainties in the adopted effective temperature, gravity and model atmospheres (e.g. Fry \\& Carney 1997). This implies that an improvement in the present result must come from an increased sample of stars with periods $\\less\\ 45 d$ spread over a range in metallicities as large as possible." }, "0403/astro-ph0403494_arXiv.txt": { "abstract": " ", "introduction": "QSOs are intrinsically luminous and therefore can be seen rather easily at large distances; but they are rare, and finding them requires surveys over large areas. As a consequence, at present, the number density of QSOs at high redshift is not well known. Recently, the Sloan Digital Sky Survey (SDSS) has produced a breakthrough, discovering QSOs up to $z=6.43$ \\cite{Fan03} and building a sample of six QSOs with $z>5.7$. The SDSS, however, has provided information only about very luminous QSOs ($M_{1450} \\mincir -26.5$), leaving unconstrained the faint end of the high-$z$ QSO Luminosity Function (LF), which is particularly important to understand the interplay between the formation of galaxies and super-massive black holes (SMBH) and to measure the QSO contribution to the UV ionizing background \\cite{Madau99}. New deep multi-wavelength surveys like the Great Observatories Origins Deep Survey (GOODS), described by M.Dickinson, M.Giavalisco and H.Ferguson in this conference, provide significant constraints on the space density of less luminous QSOs at high redshift. Here we present a search for high-$z$ QSOs, identified in the two GOODS fields on the basis of deep imaging in the optical (with {\\em HST}) and X-ray (with {\\it Chandra}), and discuss the allowed space density of QSOs in the early universe, updating the results presented in \\cite{cristiani04}. ", "conclusions": "At $z>4$ the space density of moderate luminosity ($M_{1450} \\simeq -23$) QSOs is significantly lower than the prediction of simple recipes matched to the SDSS data, such as a PLE evolution of the LF or a constant universal efficiency in the formation of SMBH in DMH. A flattening of the observed high-$z$ LF is required below the typical luminosity regime ($M_{1450} \\mincir -26.5$) probed by the SDSS. An independent indication that this flattening must occur comes from the statistics of bright lensed QSOs observed in the SDSS \\cite{Wyithe02} that would be much larger if the LF remains steep in the faint end. A similar result has been obtained at $5 \\mincir z \\mincir 6.5$, by \\cite{Barger03}. The QSO contribution to the UV background is insufficient to ionize the IGM at these redshifts. This is an indication that at these early epochs the formation or the feeding of SMBH is strongly suppressed in relatively low-mass DMH, as a consequence of feedback from star formation \\cite{granato01} and/or photoionization heating of the gas by the UV background \\cite{haiman99}, accomplishing a kind of inverse hierarchical scenario." }, "0403/hep-th0403138_arXiv.txt": { "abstract": "The effective equation of state of normal matter is changed in theories where the size of the compact space depends upon the local energy density. In particular we show how the dilution of a fluid due to the expansion of the universe can be compensated by an increase of the effective coupling of that fluid to gravity in the presence of a potential which acts to reduce the size of the compact space. We estimate how much cosmic acceleration can be obtained in such a model and comment on the difficulties faced in finding an appropriate potential. ", "introduction": "Current astrophysical observations suggest that there are at least two separate epochs in the history of the universe where accelerated expansion occurred. The first is in the very early universe where a period of accelerated expansion could explain the uniform temperature of the cosmic microwave background radiation across the sky. The second is the apparent acceleration of the universe deduced from observations of type 1a supernovae which seems to be occurring today and to have begun extremely recently in cosmological terms \\cite{sn1a}. It is not possible to obtain accelerated expansion from normal matter or radiation - as the universe expands, their energy density shrinks too rapidly, so in order to explain the observations one is forced to consider other forms of stress energy. It is very easy to obtain accelerated expansion from the potential of a self interacting scalar field, although typically one does not obtain a prolonged period of such behaviour without fine-tuning which makes it difficult to explain the inflation required in the early universe. In the same way it makes it difficult to arrange a period of acceleration which began only very recently. It is therfore worth considering other mechanisms which could give rise to acceleration to see if there are any reasonable alternatives to the orthodox mechanisms. In theories with more than 3 spatial dimensions, the values of the couplings in the 3+1 dimensional theory depend upon the details of the compactification of the higher dimensions. Perhaps the simplest example of this is the ratio between the Newton's constant which appears in the higher dimensional theory $G_{D}^{-1}=8\\pi M_F^{D-2}$ and that which appears in the 3+1 dimensional theory $G_{4}^{-1}=8\\pi M_{Pl}^2$, which is simply the volume of the compact space\\footnote{For a string theory the usual situation is $8\\pi G_{D}=M_F^{2-D}=g_s^{2}l_s^{D-2}/8$ where $g_s$ and $l_s$ are the string coupling and length respectively \\cite{witten}.}. If one then allows that volume to vary, the 4D Newton's constant will also vary relative to the underlying higher dimensional length scale. The effect of a varying volume will be to add a dynamical scalar in front of the 4D Ricci Scalar in the action ${\\mathcal{L}}=\\sqrt{-g}\\Phi R[g]+...$ but it is always possible to perform a conformal transformation on the metric to the Einstein frame such that the the action for gravity takes the form (see appendix) \\begin{equation} S=\\frac{M_{Pl}^{2}}{2}\\int d^{4}x \\sqrt{-g}\\left\\{R-\\frac{1}{2}\\partial^\\mu\\phi \\partial_\\mu \\phi\\right\\} \\end{equation} where the scalar $\\phi$ represents the volume of the compact space (variously referred to as the radion, dilaton, modulus, breathing mode etc.). If we start in $D$ dimensions and compactify to $3+1$ dimensions $\\phi$ is given by $\\cite{argurio}$ \\begin{eqnarray} \\phi(t)-\\phi_0=-\\frac{(D-4)}{2}\\sqrt{\\frac{D-2}{D-4}}\\ln\\left(\\frac{r(t)}{r_0}\\right)&&\\nonumber\\\\ \\equiv-\\frac{(D-4)}{\\beta}\\ln\\left(\\frac{r(t)}{r_0}\\right)&& \\label{beta} \\end{eqnarray} where $r_0$ is the radius of the compact (D-4) torus today so that $\\phi-\\phi_0$ parameterises the relative change in volume over time. The conformal transformation should not affect any physics derived from the Lagrangian, and indeed one finds that the effect of the varying compact space has been re-absorbed into a variation of the effective density of matter. In the Einstein frame the left hand side of the field equations will have the same form as Einstein gravity for given space-time symmetries. However, the equivalence principle is broken (the gravitational attraction between two particles will depend on the local value of the field $\\phi$) and the density that one uses to solve the equations will be given by $\\rho_{eff}=e^{\\beta\\phi}\\rho$ $\\cite{lidsey}$. If one assumes that the total density is the sum of the matter density and a potential of the (ad-hoc at this stage) form $V(\\phi)=\\bar{V}e^{-\\alpha\\phi}$ where $\\bar{V}$ and $\\alpha$ are constants, the total effective density in the Einstein frame is given by \\begin{equation} \\rho_{total}=\\bar{V}e^{-\\alpha\\phi}+\\rho e^{\\beta\\phi} \\label{mixeddensity} \\end{equation} and (for $\\omega<1/3$) the expectation value of $\\phi$ will be at the minimum of this effective density, i.e. \\begin{equation} \\langle\\phi\\rangle=\\frac{1}{\\alpha+\\beta}\\ln\\left(\\frac{4}{(1-3\\omega)}\\frac{\\alpha}{\\beta}\\frac{\\bar{V}}{\\rho}\\right)\\qquad ; \\qquad \\omega < 1/3 \\label{expectation} \\end{equation} so that although $\\phi-\\phi_0$ is always negative since $r\\ge r_0$, $\\phi$ is always positive as we will only be considering situations where $\\rho\\ll\\bar{V}$. The expectation value of $\\phi$ and the coupling of the matter to gravity therefore depends upon the local density of stress-energy. The authors of $\\cite{chameleon}$ refer to $\\phi$ as a chameleon field when it is behaving like this since the mass of the field changes according to the density of the local medium (see also \\cite{barrow}). The fact that the expectation value of $\\phi$ is a function of the local density is more important for our purposes than its mass. Consider a universe with a spatially homogeneous distribution of matter. As the universe expands, the energy density will dilute in the normal way but the expectation value of the field $\\phi$ will also change, increasing the coupling of that matter to gravity. In this way the effective dilution of the gravitating energy due to the expansion will be reduced and the effective equation of state of the energy will change. In particular we will be interested in finding out under what conditions this situation can lead to accelerated expansion. In order for such an increase in coupling to compensate for dilution, we require a potential such as the one written down in equation (\\ref{mixeddensity}) which, in the absence of matter, would cause the compact space to shrink as the universe expands. This is not easy to achieve. The reader should be aware of some closely related previous work where the mass of particles changes with the cosmically varying expectation value of a scalar field $\\cite{vamps}$ (see also \\cite{rachel}) Here, we are changing the gravitational mass of the particles whilst their inertial mass stays a fixed fraction of $M_{F}$. We are considering only cases when the mass scale associated with the energy density, $\\rho^\\frac{1}{4}$ is much less than the inverse radius of the compact space. It is therefore irrelevant as to whether the matter is confined to a brane or not, the size of the compact space will not make any difference to its bare stress energy in the Jordan-string frame since we will assume that any matter which does propagate in the bulk consists only of zero modes. The low energy gauge coupling of those fields which are able to propagate in the compact space will change with the size of the compact manifold, but we will assume that no phase transitions occur because of this effect, and will only consider particles which are given mass via some Yukawa coupling rather than some confinement scale ('electron' like particles rather than 'baryon' like ones) In the next section we will find out how the effective equation of state behaves for the simple situation presented in equation (\\ref{mixeddensity}). We will then compare the situation with that of an exponential potential and no matter, in other words power law inflation. Next we will set up some checks that our scenario must pass in order to be self consistent. Then we will find out how much expansion it is possible to obtain using matter with such a potential without violating our consistency checks. After that we will discuss why it is very difficult to find a potential like the one used above but we will show that it is possible to find some well motivated potentials which contain regions where the compact space is dynamically driven to smaller radii. Finally we will show that it is impossible to obtain acceleration using only the matter and the motion of the scalar field without another potential being present. ", "conclusions": "In this paper we showed that in models with compactified higher dimensions, changes in the size of the compact space change the effective equation of state of matter. In particular, in the presence of a potential which tends to reduce the size of the compact space, one can obtain acceleration from normal matter. Assuming a simple form for this potential we have calculated the maximum amount of expansion one could get from a such a model without non-adiabatically producing KK excitations around the compact space. If we did excite KK modes, their effect on the evolution of the compact and non-compact spaces would be extremely difficult to calculate. For a prolonged period of acceleration, the compact space would have to start out very large compared to its size at the end of this period. We have also tried to explain why it is difficult to find potentials of a suitable form to lead to acceleration, although we have shown that realistic potentials already exist which contain regions where the compact space would be pushed to smaller sizes as the matter or radiation becomes diluted. We ended by pointing out that it is hopeless to try and get acceleration without the presence of a potential. So far we have said nothing about Dark energy. It has been pointed out recently that if one is able to find an appropriate potential of the form that we have used in this paper one can give a mass to $\\phi$ which changes according to the local density $\\cite{chameleon}$. In particular, tests of tensor gravity could be passed in denser mediums such as the solar system or neutron star binary systems, while a scalar-tensor theory would describe gravity in the low density voids between clusters of galaxies. The equation of state of dark matter, could be reduced in these regions as it is in other models of varying mass particles (VAMPS) \\cite{vamps}\\cite{rachel}. In this way, one might hope to shed light on the coincidence between the energy density of dark energy and dark matter." }, "0403/astro-ph0403480_arXiv.txt": { "abstract": "We address the fine-tuning problem of dark energy cosmologies which arises when the dark energy density needs to initially lie in a narrow range in order for its present value to be consistent with observations. As recently noticed, this problem becomes particularly severe in canonical Quintessence scenarios, when trying to reproduce the behavior of a cosmological constant, i.e. when the dark energy equation of state $w_Q$ approaches $-1$: these models may be reconciled with a large basin of attraction only by requiring a rapid evolution of $w_Q$ at low reshifts, which is in conflict with the most recent estimates from type Ia Supernovae discovered by Hubble Space Telescope. Next, we focus on scalar-tensor theories of gravity, discussing the implications of a coupling between the Quintessence scalar field and the Ricci scalar (``Extended Quintessence''). We show that, even if the equation of state today is very close to $-1$, by virtue of the scalar-tensor coupling the quintessence trajectories still possess the attractive feature which allows to reach the present level of cosmic acceleration starting by a set of initial conditions which covers tens of orders of magnitude; this effect, entirely of gravitational origin, represents a new important consequence of the possible coupling between dark energy and gravity. We illustrate this effect in typical Extended Quintessence scenarios. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403355_arXiv.txt": { "abstract": "We present the results of mapping the planetary nebula NGC\\,6369 using multiple long slit spectra taken with the CTIO 1.5m telescope. We create two dimensional emission line images from our spectra, and use these to derive fluxes for 17 lines, the H$\\alpha$/H$\\beta$ extinction map, the [SII] line ratio density map, and the [NII] temperature map of the nebula. We use our photoionization code constrained by these data to determine the distance, the ionizing star characteristics, and show that a clumpy hour-glass shape is the most likely three-dimensional structure for NGC\\,6369. Note that our knowledge of the nebular structure eliminates all uncertainties associated with classical distance determinations, and our method can be applied to {\\it any spatially resolved emission line nebula}. We use the central star, nebular emission line, and optical+IR luminosities to show that NGC\\,6369 is matter bound, as about 70\\% of the Lyman continuum flux escapes. Using evolutionary tracks from \\cite{B95} we derive a central star mass of about 0.65\\,M$_{\\odot}$. ", "introduction": "The planetary nebula (PN) NGC\\,6369 -- shown in Fig.\\,\\ref{n2ima} in the light of H$\\alpha$\\,+\\,N[II]$\\lambda$658.4nm -- is an object with a complex morphology, consisting of a main bright annulus of diameter 40\\arcsec, and fainter, curved outer structures on two sides in the E-W direction. Deep narrow band H$\\alpha$\\, and \\,N[II]$\\lambda$6584 images of NGC\\,6369 were obtained by \\cite{CSSP03} in a survey to look for faint outer halos around PNe. Apart from being deeper, their image is not significantly different from the one shown in Fig.\\,1, and no large faint halo was found. The HST image of NGC\\,6369 available at http://heritage.stsci.edu/2002/25/index.html also shows the same main features as our image but has additional resolved details, and tags the H$\\alpha$, [OIII]500.7nm, and [NII]658.4nm lines with the colors red, blue and green respectively. In the section on observational results we discuss this composite HST image in more detail. \\clearpage \\begin{figure} \\includegraphics[scale=0.65]{f1.eps} \\caption{Narrow band filter image of NGC\\,6369 in H$\\alpha $+{[}NII{]} from \\cite{SCM92}. Note the bright ring and outer faint ansae to the E and W of the object. N is up, E to the left, and the plate scale is 0.26 ''/pix. \\label{n2ima}} \\end{figure} \\clearpage The equatorial and galactic coordinates for NGC\\,6369 are $\\alpha$ = 17$^h$\\,29$^m$\\,20$^s$.40 and $\\delta$ = 23$^o$\\,45\\arcmin\\,37\\arcsec.9 (ICRS2000), and l = 2.43$^o$ b = +5.85$^o$ respectively. The extreme distances determined for NGC\\,6369 are 0.33\\,kpc obtained by \\cite{AGNR84}, and 2.00\\,kpc obtained by \\cite{GPG86}, out of a total of 10 listed by \\cite{AO92}, from which we obtain a mean value for the distance to NGC\\,6369 of 1039$\\pm$ 538 pc. Note that these ten distances were determined using several statistical methods, each with their own systematic biases and assumptions, so that the average probably is a reasonably unbiased estimator, within the (large) error. The temperature of the central star obtained from the Zanstra method by \\cite{GP89} is $T(H)_\\mathrm{Z}=67\\,600~\\mathrm{K}$. The star has been classified as a WC4 star by \\cite{TAS93}. Spectra of NGC\\,6369 were obtained by \\cite{AKSJ1989}, using an IDS (Image Dissector Scanner) as detector with double 4\\,\\arcsec~apertures, by \\cite{AK87} using an ITS (Image Tube Scanner) with a $2\\arcsec\\,\\times\\,10\\arcsec$ slit, and by \\cite{PPPP} using an IDS and a combination of different apertures not bigger than $13\\arcsec\\,\\times\\,13\\arcsec$. An expansion velocity of $41.5$ km s$^{-1}$ has been determined by \\cite{MWF88}. The system radial velocity was reported to be -106\\,km.s$^{-1}$ by \\cite{W53}, and -101\\,km.s$^{-1}$ by \\cite{MWF88}. The spectral energy distribution (SED) of NGC\\,6369 is shown in Fig.\\,\\ref{sed} and has been computed using flux values from \\cite{AO92} and \\cite{SK02} over a range of wavelengths from 0.44 to 1100\\,$\\mu$m. Note that we plot $\\lambda$\\,F$_{\\lambda}$ against wavelength and below we will use the F$_{\\lambda}$ values to determine the observed optical+IR luminosity of NGC\\,6369. \\clearpage \\begin{figure} \\includegraphics[width=\\columnwidth]{f2.eps} \\caption{The SED for NGC\\,6369 between 0.44 and 1100\\,$\\mu$m using flux values from the literature (see text).\\label{sed}} \\end{figure} \\clearpage Planetary nebulae are generally classified according to their appearance, and such classification is then used for studies of the formation and evolution of these objects (for example \\citep{B87}). Such a classification, based on the observed two dimensional (2-D) brightness distribution in a given line or lines can be quite misleading. This could be the case for NGC 6369, for which a round morphology is evident from the observed images. Different 3-D geometries can produce the same observed 2-D morphology as has been shown by \\cite{M00}. Furthermore, images produced by ions of different ionization degree can be very different, due to the radiation transfer within the nebulae. Only detailed modeling, which reproduces the brightness intensity distribution of different lines, as well other observed parameters, can provide more reliable information about the 3-D geometry of the gas distribution. The determination of the actual 3-D gas distribution in planetary nebulae is essential for understanding their formation and evolution, as well as that of the ionizing star. By determining the 3-D structure of nebulae, we eliminate the large uncertainties that have plagued classical statistical distance determination methods for over 5 decades. Assumptions about the filling factor, constancy of ionized mass or diameter, mass-radius relationships etc. are not needed here: we {\\it know} what the structure and ionized mass are, and can therefore determine distances to much greater accuracy than before. In this paper we explain in detail how we can determine this 3-D structure from long slit spectra and our 3-D photo-ionization model, and apply it to the case of NGC\\,6369 In summary, we obtain the spatial structure of the object along with its chemical composition, ionizing source temperature and luminosity, mass, as well as an {\\it independent} distance, in a self-consistent manner. In \\S2 we discuss the observations and basic reduction procedures, as well as the details of the image reconstruction technique used to obtain the line intensity maps. In \\S3 the results obtained from these maps are discussed: the reddening correction of the images, total line fluxes and the computed temperature and density maps. In \\S4 we present the model results of the 3-D photoionization code, and we discuss the derived quantities Finally, in \\S5 we give our conclusions. ", "conclusions": "We have presented spectrophotometric maps of NGC\\,6369. These maps provided spatially resolved information for many emission lines and precise total fluxes for the whole nebula. The images produced with this technique were used to study the nebula with the usual diagnostic ratios. Each image was corrected for reddening pixel by pixel using the H$\\alpha $/H$\\beta$ image. This correction lead to some significant differences between the corrected and uncorrected images, as was shown in Figs.\\,5 \\& 7. The H$\\alpha $/H$\\beta$ map shows some interesting features. In Fig.\\,4 we show that the extinction is not uniform across the nebula, as it varies between 6 and 18. The structure follows the main nebular morphology, which could indicate that dust and neutral material are present. This is to be expected since the nebula shows a very clumpy structure, which may cause shadowing within the gas, allowing for the survival of grains and neutral material. The prominent features on either side of the nebula do not show significant differences in extinction when compared to the main region. The temperature map derived from the observed integrated line of sight intensities showed some structure but due to the low signal to noise ratio, this should be treated with caution. Within the errors and resolution of our observations the temperature can be considered to be constant across the nebula. We also show that the density map indicates a decrease in density for the central regions. This decrease, as has been discussed by \\cite{M00}, is not compatible with a closed shell structure. Based on this map we propose an hour-glass structure for the main nebula, which has reproduced all the observational features of NGC\\,6369. This indicates that the one needs to use the [SII] ratio in addition to images to distiguish between open and closed structures. The position of the outer condensations or ansae --which are off-set by 30$^o$ from the main nebular symmetry axis-- was obtained by matching model images with the observed line images (Fig.\\,\\ref{maps_full}) as well as matching the model [SII] density map with the observed map (Fig.\\,\\ref{dens_map}). The presence of these condensations can be accounted for by earlier ejection of matter and precession can account for their deviation from the symmetry axis. Many authors have dealt with these issues; for more complete discussions and models see for example \\cite{CFJ96}, \\cite{CFLJ95}, \\cite{G97}, and \\cite{LP97}. Some of the parameters we have derived (and use as a sanity check) are: The total luminosity of the observed lines, L$_l$\\,=\\,1150\\,L$_{\\odot}$; that derived from the SED by integrating the F$_{\\lambda}$ curve, L$_o$\\,=\\,7.1$\\,\\times$\\,10$^{-4}$\\,d$^2$ where d is the distance to NGC\\,6369 in pc, resulting in L$_o$\\,=\\,1700\\,L$_{\\odot}$ . Correcting this value using the method of Myers et al. (1987) we obtain 2550\\,L$_{\\odot}$. The luminosity of the central star is L$_{tot}$\\,=\\,8100\\,L$_{\\odot}$ so that the ratios of the line and optical+IR luminosities to the total luminosity are respectively: L$_l$/L$_{tot}$\\,=\\,0.14 and L$_o$/L$_{tot}$\\,=\\,0.3. Assuming that the absorbed flux is re-radiated in the IR, mainly the IRAS bands, the integrated optical+IR luminosity indicates that about 70\\% of the UV flux (Lyman continuum) escapes from the nebula. This is not implausible as the nebula is open at it's ``poles'' and is clumpy, allowing the radiation to pass through the many ``holes''. From the input matter distribution and abundances used in the model, we calculate the mass of the nebular ionized gas to be $M_{neb}\\,=\\,1.8\\,M_{\\odot}$. If we now use our values of luminosity and temperature for the central star and compare them with the evolutionary models of \\cite{B95} we can obtain the mass of the central core. Using this procedure we can see from Fig.\\,\\ref{evol_track} that the best fitting track for our data corresponds to a mass somewhat higher than $M_{core}\\,=\\,0.625 M_{\\odot}$, say, $\\approx\\,0.65\\,M_{\\odot}$, and to an initial mass of $M_{0}=3 M_{\\odot}$. If we sum our nebular and core masses, we obtain $M_{core}\\,+\\,M_{neb}\\,=\\,2.4\\,M_{\\odot}$ approximately. If we use the line intensity errors as a measure of the goodness of the model results and combine the luminosity and temperature errors we obtain approximately 20-25\\% for the uncertainty in $M_{core}\\,+\\,M_{neb}$, placing the obtained initial stellar mass within the uncertainty of our determination. The value for $M_{neb}$ calculated from the model input density is a lower limit estimate of the total nebular mass as at least some material is present in the form of dust, as indicated qualitatively by our extinction map structure. Additionally, there may be neutral gas for wich we have no constraints. Note that typical parameters for a WC4 PN central star from \\cite{A64} are T\\,=\\,90kK, and R\\,=\\,0.4\\,R$_{\\odot}$, very close to our derived values of 91000\\,K and 0.4\\,R$_{\\odot}$. Using g\\,=\\,G\\,.\\,M\\,/\\,R$^2$ in CGS units, we obtain log(g)\\,=\\,5.1, which is much lower than the value of 5.8 derived from the depth of our He break. This is due to the fact that the central star has an extended atmosphere, again showing the consistency of the model results. \\clearpage \\begin{figure} \\includegraphics[width=\\columnwidth]{f12.eps} \\caption{Comparison of model temperature and luminosity obtained for the central star with model tracks calculated by \\cite{B95}.\\label{evol_track}} \\end{figure} \\clearpage Using our photoionization code and the proposed structure we obtained a complete 3-D model for the NGC\\,6369. The fitted model line intensities show excellent agreement with the observed values. The obtained distance of d=1550 pc, is well within the range present in the literature obtained from different methods. The model temperature for the ionizing star is similar to the Zanstra (He) value discussed in \\S3.3. The temperature and luminosity values obtained from the model as well as the total nebular plus central star mass show good agreement with the stellar evolution models of \\cite{B95}. Using multiple long slit spectroscopy, we can determine accurate distances, 3-D structures, abundances, ionized masses, central star masses, luminosities, and temperatures to any spatially resolved emission line nebula, assuming there are no strong shocks or extreme morphologies involved." }, "0403/astro-ph0403025_arXiv.txt": { "abstract": "{We report the first likely spectroscopic confirmation of a $z\\sim 10.0$ galaxy from our ongoing search for distant galaxies with ISAAC/VLT. Galaxy candidates at $z \\gtapprox 7$ are selected from ultra-deep $JHKs$ images in the core of gravitational lensing clusters for which deep optical imaging is also available, including HST data. The object reported here, found behind Abell 1835, exhibits a faint emission line detected in the $J$ band, leading to $z=10.0$ when identified as \\lya, in excellent agreement with the photometric redshift determination. Redshifts $z < 7$ are very unlikely for various reasons we discuss. The object is located on the critical lines corresponding to z$=$9 to 11. The magnification factor $\\mu$ ranges from 25 to 100. For this object we estimate $SFR(\\lya) \\sim (0.8-2.2)$ \\msun yr$^{-1}$ and $SFR(UV) \\sim (47-75)$ \\msun\\ yr$^{-1}$, both uncorrected for lensing. The steep UV slope indicates a young object with negligible dust extinction. SED fits with young low-metallicity stellar population models yield (adopting $\\mu=25$) a lensing corrected stellar mass of $M_\\star \\sim 8 \\times 10^6$ \\msun, and luminosities of $2 \\times 10^{10}$ \\lsun, corresponding to a dark matter halo of a mass of typically $M_{\\rm tot} \\ga 5 \\times 10^8$ \\msun. In general our observations show that under excellent conditions and using strong gravitational lensing direct observations of galaxies close to the ``dark ages'' are feasible with ground-based 8-10m class telescopes. ", "introduction": "Spectacular progress during the last decade has permitted direct observations of galaxies and quasars out to redshifts $z \\sim 6.6$ (Hu et al.\\ 2002, Fan et al.\\ 2003, Kodaira et al.\\ 2003), or in other words over more than 90 \\% of cosmic time. With the exception of the cosmic microwave background, direct exploration of more distant objects has so far been hampered by the need for observations beyond the optical domain, and technical difficulties related to near-IR observations of such faint objects. However, by combining a strong magnification of background galaxies by well known gravitational lensing clusters with present-day near-IR instruments on 8-10m class telescopes such an attempt appears feasible (Schaerer \\& Pell\\'o\\ 2002; Pell\\'o et al.\\ 2003). We present here the first spectroscopic results obtained from our prototype observational program with ISAAC/VLT on one of our best $z \\ga 7$ galaxy candidates. In Sect.~\\ref{observations} we summarize the observational strategy adopted in the present study, and we present the photometric and spectroscopic characteristics of this object. The redshift identification is discussed in Sect.\\ \\ref{s_z}. The physical properties of this galaxy and implications of our finding are briefly discussed in Sect.~\\ref{results}. ", "conclusions": "" }, "0403/astro-ph0403096_arXiv.txt": { "abstract": "The study of the universe at energies above 100 GeV is a relatively new and exciting field. The current generation of pointed instruments have detected TeV gamma rays from at least 10 sources and the next generation of detectors promises a large increase in sensitivity. We have also seen the development of a new type of all-sky monitor in this energy regime based on water Cherenkov technology (Milagro). To fully understand the universe at these extreme energies requires a highly sensitive detector capable of continuously monitoring the entire overhead sky. Such an instrument could observe prompt emission from gamma-ray bursts and probe the limits of Lorentz invariance at high energies. With sufficient sensitivity it could detect short transients ($\\sim$15 minutes) from active galaxies and study the time structure of flares at energies unattainable to space-based instruments. Unlike pointed instruments a wide-field instrument can make an unbiased study of all active galaxies and enable many multi-wavelength campaigns to study these objects. This paper describes the design and performance of a next generation water Cherenkov detector. To attain a low energy threshold and have high sensitivity the detector should be located at high altitude ($>$ 4km) and have a large area ($\\sim$40,000 m$^2$). Such an instrument could detect gamma ray bursts out to a redshift of 1, observe flares from active galaxies as short as 15 minutes in duration, and survey the overhead sky at a level of 50 mCrab in one year. ", "introduction": "\\label{sec:Introduction} The past 15 years have seen large advances in the capabilities of ground-based gamma ray detection, from the pioneering observation of the Crab Nebula by the Whipple observatory in 1989\\cite{weekes1989} to the new generation of air Cherenkov telescope arrays such as HESS\\cite{hofmann2003}, VERITAS\\cite{veritas}, and CANGAROO\\cite{yoshikoski1999} and large area air Cherenkov telescopes such as STACEE \\cite{hanna2002}, CELESTE\\cite{pare2002}, and MAGIC\\cite{lorenz2003}. There are now at least 10 known sources of very-high-energy (VHE) gamma rays\\cite{horan2003}. The physics of these objects is astounding: from spinning neutron stars to super-massive black holes, these objects manage to accelerate particles to energies well in excess of 10 TeV. How this acceleration occurs is not well understood and there is not universal agreement on what particles are being accelerated in some of these sources. At lower energies EGRET has detected over 270 sources of high-energy gamma rays\\cite{hartman1999} and GLAST is expected to detect several thousand sources. In addition there are transient sources such as gamma-ray bursts that have to date eluded conclusive detection in the VHE regime (despite some tantalizing hints\\cite{atkins2000}). The paucity of VHE sources can be traced to the nature of the existing instruments: they are either narrow field instruments that can only view a small region of the sky at any one time and can only operate on clear moonless nights (Whipple, HEGRA, etc.) or large field instruments with limited sensitivity (Milagro, Tibet Array). The Milagro observatory has pioneered the use of a large area water Cherenkov detector for the detection of extensive air showers. Since an extensive air shower (EAS) array directly detects the particles that survive to ground level it can operate continuously and simultaneously view the entire overhead sky. With the observation of the Crab Nebula and the active galaxies Mrk 421 and Mrk 501, Milagro has proven the efficacy of the technique and its ability to reject the cosmic-ray background at a level sufficient to detect sources\\cite{atkins2003}. At the same time the Tibet group\\cite{amenomori1999} has demonstrated the importance of a high-altitude site and what can be accomplished with a classical scintillator array at extreme altitudes. A detector with the all-sky and high-duty factor capabilities of Milagro, but with a substantially lower energy threshold and a greatly increased sensitivity, would dramatically improve our knowledge of the VHE universe. Reasonable design goals for such an instrument are: \\begin{itemize}% \\item Ability to detect gamma-ray bursts to a redshift of 1.0 \\item Ability to detect AGN to a redshift beyond 0.3 \\item Ability to resolve AGN flares at the intensities and durations observed by the current generation of ACTs \\item Ability to detect the Crab Nebula in a single transit \\end{itemize} This paper describes a design for a next generation all-sky VHE gamma-ray telescope, the HAWC (High Altitude Water Cherenkov) array, that satisfies these requirements. To quantify the definition of observing ``short'' flares from AGN, previous measurements of flare intensities and durations by air Cherenkov telescopes can be used. To date the shortest observed flares have had $\\sim$15 minute durations with an intensity of 3-4 times that of the Crab\\cite{gaidos1996}. The low energy threshold needed to accomplish these goals requires that the detector be placed at extreme altitudes (HAWC would be situated at an altitude of $\\sim$4500 meters) and the required sensitivity demands a large area detector - of order 40,000 m$^2$. Section \\ref{sec:Particle_Detection} discusses the limiting performance of an EAS array based on the properties of the EAS, section \\ref{sec:Detector_Description} gives a physical description of the HAWC and section \\ref{sec:Detector_Performance} details the expected performance of HAWC. ", "conclusions": "\\label{sec:Conclusions} A design for the next generation all-sky VHE gamma-ray telescope has been presented. This instrument, HAWC, would be over 20 times more sensitive than the Milagro detector. With the ability to continuously view the entire overhead sky HAWC will be an excellent complement to both GLAST and the coming generation of air Cherenkov telescopes (HESS, MAGIC, VERITAS, and CANGAROO III). With comparable sensitivity to GLAST it will be the only instrument capable of monitoring the many thousands of sources that GLAST is expected to detect at higher energies. In addition to searching the sky for galactic sources (the VHE complement to the 150 EGRET unidentified objects), and active galaxies, the low-energy sensitivity of a detector placed at high altitude ensures that such an instrument will detect any VHE emission from gamma-ray bursts. Perhaps most importantly an open aperture instrument with this level of sensitivity could discover completely new and unexpected phenomena that have so far eluded detection." }, "0403/astro-ph0403575_arXiv.txt": { "abstract": "s{ We report results from {\\it XMM-Newton} observations of thirteen X-ray bright BL Lacertae objects, selected from the {\\it Einstein} Slew Survey sample. The spectra are generally well fit by power-law models, with four objects having hard ($\\alpha<1; F_\\nu \\propto \\nu^{-\\alpha}$) spectra that indicates synchrotron peaks at $>5$ keV. None of our spectra show line features, indicating that soft X-ray absorption ``notches'' must be rare amongst BL Lacs, rather than common or ubiquitous as had previously been asserted. We find significant curvature in most of the spectra. This curvature is almost certainly intrinsic, as it appears nearly constant from 0.5 to 6 keV, an observation which is inconsistent with the small columns seen in these sources.} ", "introduction": "The nature of the X-ray emission and absorption from BL Lacs is still an open question. Most BL Lac objects have X-ray spectra that can be fit by power-laws within smaller bandpasses, such as that of {\\it ROSAT} ([12],[17]) More recent results from {\\it ASCA} [7] and {\\it BeppoSAX} ([1],[11],[19]) have generally confirmed this spectral morphology, and also added onto it the possibility of intrinisic spectral curvature across a wider bandpass ([4],[6],[9],[16]). In addition, earlier missions had indicated that some BL Lac objects showed a deficit in soft X-rays below a power-law model, which had been interpreted by invoking X-ray absorption features at 0.5--0.8 keV ([3],[8],[14],[15]). But not all bright BL Lacs were found to require such features (e.g., [5]). ", "conclusions": "" }, "0403/astro-ph0403269_arXiv.txt": { "abstract": "We present mass and radius derivations for a sample of very young, mid- to late M, low-mass stellar and substellar objects in Upper Scorpius and Taurus. In a previous paper, we determined effective temperatures and surface gravities for these targets, from an analysis of their high-resolution optical spectra and comparisons to the latest synthetic spectra. We now derive extinctions, radii, masses and luminosities by combining our previous results with observed photometry, surface fluxes from the synthetic spectra and the known cluster distances. These are the first mass and radius estimates for young, very low mass bodies that are {\\it independent} of theoretical evolutionary models (though our estimates do depend on spectral modeling). We find that for most of our sample, our derived mass-radius and mass-luminosity relationships are in very good agreement with the theoretical predictions. However, our results diverge from the evolutionary model values for the coolest, lowest-mass targets: our inferred radii and luminosities are significantly larger than predicted for these objects at the likely cluster ages, causing them to appear much younger than expected. We suggest that uncertainties in the evolutionary models - e.g., in the choice of initial conditions and/or treatment of interior convection - may be responsible for this discrepancy. Finally, two of our late-M objects (USco 128 and 130) appear to have masses close to the deuterium-fusion boundary ($\\sim$9--14 Jupiters, within a factor of 2). This conclusion is primarily a consequence of their considerable faintness compared to other targets with similar extinction, spectral type, and temperature (difference of $\\sim$ 1 mag). Our result suggests that the faintest young late-M or cooler objects may be significantly lower in mass than current theoretical tracks indicate. ", "introduction": "The last few years have witnessed a dramatic swelling in the ranks of objects at the bottom of the Main Sequence, and in the substellar regime beyond. Hundreds of ultra-low-mass stars and brown dwarfs have been uncovered, both in the field and in star-forming regions. Studies of young clusters even suggest the presence of isolated planetary mass objects \\citep{Zapa00, Lucas00}. The existence and properties of all these low-mass bodies have profound implications for a host of issues, ranging from the dominant mechanisms for star and planet formation \\citep{Boss01, Reipurth01, Bate02, Padoan02}, to the birthline and early evolution of low-mass objects \\citep{Hartmann03}, to the shape of the initial mass function \\citep{Briceno02}. A reliable determination of mass is intrinsic to the ultimate resolution of these questions. Presently, masses (and ages) are most widely inferred by comparing observables such as temperature and luminosity to the predictions of theoretical evolutionary tracks. However, these models remain largely unverified for very low masses. The simplest test is to derive dynamical masses for the components of binary (or higher-order) systems with known orbital parameters, and compare them to the theoretical values derived from other, directly observed quantities (e.g., $\\lbol$ and \\teff). Unfortunately, this is impeded for very low-mass stars and substellar objects by the current paucity of suitable multiple systems. In most known cases, one can either deduce dynamical masses but not theoretical ones (because the components are not directly detected), or vice versa (because the orbital parameters remain indeterminate). The one exception is HD 209458, in which both are available \\citep{Charbonneau02}. The comparison of theory to observations in this case does reveal some large uncertainties in the former, and underlines the usefulness of such tests (Baraffe et al. 2003; Burrows, Sudarsky \\& Hubbard 2003). However, it is not very illuminating as a general evaluation of the models: the proximity between the planetary companion and the star in this instance engenders special insolation effects, precluding an extension of the results to free-floating brown dwarfs and planetary mass objects (or to planets with larger orbital radii). The situation is likely to improve in the near future, at least in the field - several promising systems with directly detected, probably substellar components have now come to light; dynamical masses should be obtained fairly soon \\citep{Close02, Potter02, Lane01}, allowing checks on the theoretical models for field brown dwarfs. In young clusters and star-forming regions, however, no suitable systems have emerged yet. This is especially troubling since even the identification of objects as substellar currently depends, at these early ages, on the theoretical tracks (empirical tests of substellarity that depend on Lithium detection or minimum Main Sequence temperature are largely inapplicable to very young objects). Moreover, the low-mass tracks are most uncertain precisely at such early times \\citep{Baraffe02}, so testing them for young objects is particularly crucial. To address this issue, we have developed a technique for calculating masses for young cluster objects from surface gravity measurements, independent of theoretical evolutionary models. The essential idea is simple: derive surface gravity and effective temperature (\\teff) by comparing the observed spectrum to the latest synthetic ones, then derive radius (and extinction) by combining the observed photometry and known cluster distance with the surface fluxes predicted by the synthetic spectra (for the inferred \\teff and gravity), and finally derive mass by combining radius and gravity. The sticking point, of course, is the derivation of sufficiently accurate surface gravities from the spectra, which has long been one of the major goals in the study of ultra-low mass objects. However, we have shown in a previous paper (Mohanty et al. 2003a; henceforth Paper I) that the current generation of highly detailed synthetic spectra is equal to the task. Employing these, we have derived gravities to within $\\pm$ 0.25 dex (and \\teff to within $\\pm$ 50K) in a sample of very low-mass objects in the Upper Scorpius and Taurus clusters (Paper I). We now derive masses and radii for these, using our Paper I results together with photometry and distance estimates. We will show that our analysis allows mass to be determined to within a factor of $\\sim$ 2, and radius to within $\\sim$ 30\\%. These errors are much larger than those associated, for example, with dynamical mass and radius measurements in eclipsing binaries. Nevertheless, we will demonstrate that they are sufficient for first order tests of the theoretical evolutionary tracks. Though our analysis is independent of the evolutionary models (and thus serves as a check on the latter), it is clearly dependent on the validity of the synthetic spectra we use. The accuracy of these were discussed in Paper I, and will be addressed further in this work. However, we point out that our derivation of physical parameters using spectral synthesis alone does not constitute a great leap of faith, any more than employing evolutionary models for this purpose does. There are two reasons for this. First, the $P$-$T$ structure of the deep atmosphere (which forms the inner boundary of the spectral calculations) acts as the outer boundary condition of the interior calculations; i.e., the evolutionary models are anchored with the same (deep) atmospheric modeling as the synthetic spectra. Second, in order to compare observations to the evolutionary predictions, synthetic spectra are crucial: either to convert an observed spectral type to \\teff (when placing objects on a theoretical \\teff-luminosity H-R diagram), or to convert predicted effective temperatures and luminosities to photometric colors and magnitudes (when placing objects on a theoretical color-magnitude diagram). We have only taken the dependence on synthetic spectra a step further, by using them to derive surface gravities as well; the accuracy of our gravities is discussed at length in Paper I. The advantage of this route lies in our avoiding (and thereby testing) what are perhaps the greatest uncertainties in the theoretical models for young objects: the initial conditions and still-relevant effects of accretion and collapse during the formation stage. The main disadvantage of eschewing the evolutionary models is that we cannot independently estimate ages. This is compensated for by our ability to independently estimate the mass, and thus provide a check on the evolutionary model predictions for a group of objects that belong to the same cluster, and are thus likely to be (nearly) coeval. Finally, we elucidate the system of nomenclature we have adopted here for very low-mass objects. There are considerable differences within the community at present regarding this issue: naming conventions based on fusion (or equivalently, mass), origins and location (isolated or in orbit around a star) have all been proposed. Controversies arise because the different definitions do not yield the same grouping of objects; which of these systems is finally adopted is a matter for future arbitration. However, given that a consensus is currently lacking, and that our primary concern in this paper is mass, we adopt a fusion-based convention (since mass is most directly associated with the presence and type of fusion). The term `brown dwarf' refers to all objects which never derive 100\\% of their luminosity from hydrogen burning (unlike stars), but which are nevertheless above the deuterium-burning mass limit. Thus brown dwarfs are objects in the range $\\sim$ 0.012--0.080 \\msun (12--80 \\mj). We contract the term `planetary mass object' to the less cumbersome `planemo', and use it to refer to all objects below the deuterium-burning limit (i.e., mass $\\lesssim$12 \\mj), regardless of whether they are free-floating or in orbit around a star. When a distinction is required between the two cases (e.g., when referring specifically to the recent isolated planemo candidates), it will be made explicitly. Since neither planemos nor brown dwarfs undergo stable hydrogen fusion (i.e., reach the Main Sequence), both are included under the rubric of `substellar objects'. Since both stars and brown dwarfs undergo at least some fusion, they will collectively be termed `fusors'; in this context, planemos are non-fusors. A broader discussion of these terms and nomenclature issues is presented in Basri 2003. ", "conclusions": " {\\it (1)} Both radius and \\teff decrease less rapidly with diminishing mass, at a given young age, than predicted by the theoretical evolutionary models. Specifically, in the mass-radius plane the lowest mass objects ($\\lesssim$30 M$_J$) remain much larger (i.e., contract more slowly with age) than the models suggest, while the higher masses have radii in good agreement with the model predictions. In the mass-\\teff plane, the higher masses are substantially cooler than predicted, while the lowest masses have \\teff either in better agreement with, or hotter than, the model values. The combination of these two trends implies that luminosity also falls off less dramatically with mass, at a given age, than the evolutionary models indicate. {\\it (2)} The lowest masses in our Upper Sco sample are near the deuterium fusion boundary. Because of the importance of both conclusions, we have taken considerable pains to consider possible sources of error, both observational and systematic. These include conversion of colors to extinctions, temperature scales for pre-main sequence objects, problems with the gravity measurements, and the effects of starspots or binarity. Our extinctions are consistent with an analysis of the same region using low dispersion spectra. Our temperature scale is in good agreement with recent photometric work in the field. The range of masses we find within a few spectral subclasses is perhaps surprising, but we show that some of our basic conclusions can be drawn just from the observations (without recourse to theory at all). We conduct a comparative analysis with a more extensively-studied young (GG Tau B) binary system to further test our conclusions, and find comparable discrepancies with theory in that case as well. Finally, our derived relationships between radius, mass, \\teff and luminosity all agree (within the measurement uncertainties) with certain basic theoretical predictions that are likely to be correct regardless of evolutionary model uncertainties: younger objects have larger radii than older ones of the same mass; less massive objects are cooler and generally smaller than more massive ones at a given age; and luminosity decreases with both diminishing mass (at fixed age) and increasing age (at fixed mass). There does not appear to be any a priori physical basis, therefore, for discarding our results. We also point out that (like the Lyon models) our mass-radius, mass-\\teff, radius-\\teff and mass-luminosity relationships are smooth, without any sharp breaks or discontinuities. The precipitous drop in gravity at low temperatures that we found in Paper I, which might seem remarkable at first sight, is due (if our analysis is correct) simply to a relatively slow change in radius and \\teff (compared to the Lyon predictions) over a significant range in mass. The weight of the evidence suggests that substantially more work should go into the measurement of physical parameters of young substellar objects, the validity of the evolutionary tracks, and, without doubt, further testing and confirmation of our results. An especially important conclusion of our work is that agreement with the evolutionary models in any single two-parameter plane (e.g., mass-radius) does not guarantee agreement in all parameters (e.g., \\teff, luminosity). In order to ascertain the veracity of the models, their predictions must be checked for all the parameters, not just a selected few. As a corollary, comparing an object to the evolutionary models over one set of parameters (e.g., \\teff-luminosity), in order to estimate other quantities (e.g., mass), is an exercise that is not always justified. Such translations, which are common practice in current studies of young low-mass objects, may lead to spurious mass and radius estimates, and must be undertaken with great caution. Similar conclusions have been reached by other authors, in the context of evolutionary model comparisons to higher-mass (solar-type) PMS stars (e.g., Torres \\& Ribas 2002). In Paper I, we pointed out some specific areas of concern for theory, such as accretion effects and the treatment of convection and deuterium fusion. In particular, we noted that if deuterium fusion begin at an earlier time than predicted, the discrepancies in radius and gravity between the theoretical tracks and our measurements, for the lowest masses, may be resolved. This is a testable hypothesis, as we outlined in Paper I, and bears closer examination. In this paper, we have also identified discrepancies in the theoretical \\teff predictions (assuming our derived temperatures are accurate), especially for the higher mass objects in our sample. The underlying physical basis for temperature uncertainties in the evolutionary models is unclear; it is possible that remaining inadequacies in the treatment of convection are at fault. Finally, while the model atmospheres and synthetic spectra that lie at the heart of our analysis are tremendously improved from earlier generations, they still suffer from certain shortcomings. Specifically, they reproduce the photometry of field M dwarfs in some, but not all, of the optical and infrared bands. While we have gone to great lengths to account for, and exclude, any attendant uncertainties in our analysis, further improvements in the atmospheric modeling - particularly in the linelists and opacities (most importantly, of H$_2$O) - would be tremendously useful for future studies of field and PMS low-mass objects. We have implemented methods that have long been used for normal stars. They provide a means of testing theoretical isochrones and obtaining fundamental stellar parameters for very young, very low-mass objects. This methodology (which highlights the importance of high resolution spectroscopy and model atmospheres) should also be extended to higher mass objects and other star-forming regions with different ages. Extensive programs of this nature are now both desirable and feasible." }, "0403/astro-ph0403119_arXiv.txt": { "abstract": "{ We present simultaneous multifrequency radio observations for a complete subsample of 26 XBLs from the {\\it Einstein} Extended Medium-Sensitivity Survey, obtained with the Very Large Array (VLA). Spectra are computed using fluxes at 20, 6 and 3.6 cm. Unlike many radio selected samples, the EMSS did not impose any criterion on the radio spectrum to identify BL Lac objects. It is therefore possible to investigate the intrinsic radio spectral slope distribution and to determine the effect produced by this selection criterion. We find that 15\\% of the observed objects do not meet the flat-spectrum criterion imposed on some other BL Lac samples. A dataset that includes non-simultaneous data (that are also taken with different VLA configurations) shows an even higher percentage of steep spectrum sources. This effect can be ascribed to a larger fraction of extended flux detected with the more compact VLA configuration. Possible biases introduced by the flat--radio-spectrum criterion in the radio-selected BL Lac samples cannot explain the discrepancies observed in the evolutionary properties of Radio and X-ray selected samples of BL Lacs. ", "introduction": "BL Lac objects are an enigmatic class of active galactic nuclei (AGN) characterized by strong radio, optical and X-ray variability, relatively high optical and radio polarization and featureless optical spectra (e.g. Urry \\& Padovani, 1995). The properties of the class can be explained in terms of the {\\it{relativistic beaming}} scenario wherein the observed emission is dominated by Doppler-boosted non-thermal radiation from a relativistic jet aligned with the line of sight (Blandford \\& Rees, 1978, Antonucci \\& Ulvestad, 1985). Owing to their featureless spectra and lack of UV excess in many BL Lacs, optical search techniques (e.g. by excess colors, emission line strength), used to search for other AGN, failed to find BL Lacs in large number (e.g. Fleming et al., 1993). Since BL Lacs are both radio-loud and X-ray-loud, surveys in these frequency bands discover them with high efficiency (e.g. the ``1 Jansky'' sample: Stickel et al., 1991; the {\\it Einstein} Extended Medium-Sensitivity Survey (EMSS) sample: Morris et al., 1991; Rector et al., 2000; the Deep X-ray Radio Blazar Survey (DXRBS) sample: Perlman et al., 1998, Landt et al. 2001; the ROSAT All Sky Survey (RASS) sample: Bade et al., 1998; the ``sedentary'' survey: Giommi, Menna \\& Padovani, 1999; the Radio Emitting X-ray (REX) survey sample: Caccianiga et al., 1999, Caccianiga et al., 2002). This splits BL Lacs into two empirical subclasses, namely {\\it radio-selected BL Lacs} (RBLs) and {\\it X-ray-selected BL Lacs} (XBLs). Padovani \\& Giommi (1995) proposed to distinguish these two classes on a more physical basis, according to the spectral energy distribution (SED). For ease of description here we will refer to RBL and XBL only. RBLs and XBLs show different behaviors with respect to properties like cosmological evolution, polarization, variability, core-to-extended radio flux ratio, extended radio morphology, spectral features (e.g. % Wolter et al., 1994; Stickel et al., 1991 Morris et al., 1991; Jannuzi, Smith \\& Elston, 1993, 1994; Perlman \\& Stocke, 1993 to mention just a few) which have not been satisfactorily explained yet. This led to the suggestion that the observed discrepancies are at least partially caused by selection effects due to different criteria used in radio and X-ray surveys. In particular, the standard selection technique for RBLs requires a flat-radio-spectrum criterion for which the radio spectral slope must be less than 0.5 ($\\alpha_r\\leq0.5$; $S_\\nu \\propto\\nu^{-\\alpha_r}$). This criterion has not been used for classical X-ray selected samples of BL Lacs like the EMSS. If surveys with this criterion select only part of the BL Lac population, the RBL samples could be biased and the differences between RBLs and XBLs like the ones mentioned above could be at least partially explained. It is already known that a number of objects with steep spectra ($\\alpha_r>0.5$), classified as radio galaxies in radio surveys like the ``1 Jansky'' (Owen, Ledlow, \\& Keel, 1996, Perlman et al. 1996), otherwise meet the BL Lacs selection criteria, based on optical properties (equivalent width of emission lines; Ca~II break contrast) of March\\~a et al. (1996), and have broadband properties (radio luminosity and overall spectral energy distributions) that agree with those of BL Lacs (e.g. Rector et al., 2000). To test if the flat--radio-spectrum criterion really introduces biases in the BL Lacs selection and to estimate the incompleteness degree of RBLs samples, we studied the radio spectral indices of a complete subsample of XBLs in which the identification criteria were not based on the radio spectral properties. Previous multi-frequency studies had already been performed on XBLs. For instance, Laurent-Muehleisen et al. (1993) collected non-\\-simultaneous data for several objects of this class, that however do not make a complete sample. We prefer to deal with the largest subsample for which we could have both simultaneous and non-simultaneous data. The sources of this subsample are extracted from the EMSS (Gioia et al., 1990; Stocke et al., 1991), a % catalog whose BL Lacs sample has been thoroughly studied for completeness (Rector, Stocke \\& Perlman, 1999). A pilot study on the radio spectra of 8 XBLs from the EMSS and the {\\it HEAO-1 A-2} all-sky survey (Piccinotti et al., 1982) was performed by Stocke et al. (1985). Three objects showed a spectral slope which exceeds the limit for flat spectrum and two more were marginally steep. Preliminary analysis of the EMSS BL Lacs sample, by using non-simultaneous data at 6 and 20 cm from the EMSS and the {\\it NRAO VLA Sky Survey} (NVSS: Condon et al., 1998) respectively, showed that about 30\\% of the objects have a spectral slope steeper than 0.5. This is the motivation that led us to investigate the EMSS sources by obtaining simultaneous spectra, which are not affected by variability. We briefly describe the sample of objects used for this work in \\S\\ref{The Sample}; in \\S\\ref{Data analysis} we present the data reduction process as well as flux densities and radio spectra for each object; in \\S\\ref{Results} we show the results obtained; our conclusions are summarized in \\S\\ref{Conclusions}. Throughout the paper we used $H_0=50$ km s$^{-1}$ Mpc$^{-1}$ and $q_0=0$, but no conclusions are dependent upon this choice of cosmology. ", "conclusions": "\\label{Conclusions} By using fluxes at 20, 6 and 3.6 cm, simultaneously measured with the VLA, we computed the radio spectra for a complete subsample of 22 XBLs from the EMSS (Gioia et al., 1990, Stocke et al., 1991). To this sample we added 4 sources whose simultaneous spectra were obtained by Stocke et al. (1985). The aim was to study the spectral slope distribution at radio frequencies without possible biases introduced by the flat-radio-spectrum criterion often used in selecting BL Lac samples. We found that about 15\\% of the sources have steep spectra. We considered the archived non-simultaneous data as well, finding an even higher percentage of steep spectra. This effect could be ascribed to variability, but other factors seem more significant in determining the measured slope of spectra. The higher percentage of steep spectra objects ($\\sim$ 38\\%), obtained by using the 20cm NVSS observations, implies the presence of a large fraction of extended flux that is preferentially detected by the VLA D configuration. This seems also supported by a possible correlation between spectral index and 20cm extended luminosities, computed either with simultaneous or non-\\-simultaneous data. We find that the possible biases introduced by the flat--radio-spectrum criterion in the RBLs samples cannot easily account for the discrepancies observed in the $\\langle V_e/V_a \\rangle$ values of RBLs and XBLs samples and therefore the issue of evolution of different BL Lac classes is still open. The percentages of steep spectrum BL Lacs found for the EMSS sample cannot be applied straightforwardly to RBL samples, since a) the selection by radio spectral index is sometimes computed using non-simultaneous data; b) the XBLs properties are different from those of RBLs in the sense that RBLs are more variable and possibly more core dominated. If the observed spread in $\\alpha_r$ is mostly due to variability, we should expect a larger fraction of steep RBLs than XBLs. If most of the effect is due, as the data seem to suggest, to the presence of a significant extended component, that is lesser in RBL, the influence on the radio selected objects should be less severe. We stress here that the $\\alpha_r$ distribution in the EMSS BL Lacs simultaneous data is a continuous one and so any division in two subsamples is somewhat arbitrary and does not have a physical underlying support. We are tempted to suggest a different cut-off value of $\\alpha_r=0.7$ (as done e.g. by Perlman et al. 1998 in DXRBS) that would then exclude $\\leq$ 10\\% of all BL Lacs based upon our findings here. But, because this would also include a very large number of ``normal'' radio galaxies, some of which with optical featureless spectra (see e.g. discussions in Rector \\& Stocke, 2001, Perlman et al. 1996), other information would have to be used to define them as BL Lacs (e.g. variable optical polarization). However, we have shown that the selection effects applied by using a cut-off value in the radio spectral index of $\\alpha_r=0.5$ are not so severe. The steeper BL Lacs that are excluded from this criterion seem to have the same overall properties of the flat ones, especially for what concerns the cosmological distribution. We can therefore conclude that the discrepancies found by using different samples cannot be ascribed to the flat radio spectrum criterion." }, "0403/astro-ph0403433_arXiv.txt": { "abstract": "Given the scientific goals of VSOP-2, including the possibility of observations of the shadows of black holes, we have investigated the fidelity of the recovered images given a typical {\\em uv}-coverage. We find that we can achieve a dynamic range of better than 1000:1. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403009_arXiv.txt": { "abstract": "{With the advent of the present and future spatial X-ray missions, it becomes crucial to model correctly the line spectrum of X-ray emitting media such as the photoionized plasma observed in the central regions of Active Galactic Nuclei (AGN), or in X-ray binaries. We have built a photoionization code, Titan, solving the transfer of a thousand lines and of the continuum with the ``Accelerated Lambda Iteration\" method, which is one of the most efficient and at the same time the most reliable for line transfer. In all other photoionization codes the line intensities are computed with the so-called ``escape probability formalism\", used in its simplest approximation. In a previous paper (Dumont et al. 2003), it was shown that this approximation leads to a wrong estimation of the emitted X-ray line intensities, especially in the soft X-ray range. The errors can exceed one order of magnitude in the case of thick media (Thomson thickness of the order of unity). In the present paper, we show that it also happens, but for different reasons, in the case of moderately thin media (Thomson thickness of 0.001 to 0.1), characteristic of the Warm Absorber in Seyfert 1 or of the X-ray emitting medium in Seyfert 2. Typically, the errors on the line fluxes and line ratios are of the order of 30$\\%$ for a column density of 10$^{20}$ cm$^{-2}$, and a factor five for a column density of 10$^{23}$ cm$^{-2}$, in conditions giving rise to the spectra observed in these objects. We explain why this problem is less acute in cooler media, like the Broad Line Region of AGN. We show some examples of X-ray spectra appropriate for Seyfert 2 and for the Warm Absorber of Seyfert 1. We conclude that though it is quite important to introduce numerous accurate X-ray data in photoionization codes, it should be accompanied by more elaborate methods than escape probability approximations to solve the line transfer. ", "introduction": "Since more than three decades, photoionization codes have been developed to compute the structure and the spectrum of photoionized media, such as HII regions, planetary nebulae, supernova remnants, envelopes of novae, Narrow Line Regions of Active Galactic Nuclei (AGN), etc\\ldots At the end of the seventies, these codes have been extended to denser and thicker media, like the Broad Emission Line Region (BLR) in quasars and AGN or the emission regions of X-ray binaries and cataclysmic variables. For this purpose, the formalism of ``escape probability\" has been introduced to take into account self-absorption in lines while avoiding to solve the line transfer (Netzer 1975), and a new type of photoionization code has begun to be built. These codes - Cloudy (Ferland et al. 1988), XSTAR (Kallman \\& Krolik 1995), ION (Netzer 1993), for instance - have now reached a high degree of sophistication, including very accurate atomic data, a large number of ions and transitions, and all the necessary processes allowing to use them in various physical conditions. With the advent of the X-ray missions Chandra and XMM Newton, splendid spectra of various types of objects have been obtained in the soft X-ray range, showing tens of emission lines which can be used as diagnostics of the physical state of the emitting region. The best examples are Seyfert 2 galaxies, which display a rich X-ray line spectrum, most probably produced by the external part of the ``Warm Absorber\" of Seyfert 1, photoionized by the intense central continuum and seen in emission because the central continuum is hidden from our view (Antonucci \\& Miller 1985). Typical column densities of this medium are 10$^{21-23}$ cm$^{-2}$ (Sako et al. 2000, Kinkhabwala et al. 2002, Ogle et al. 2003). Most naturally, the photoionized codes developed for the BLR have been used to model the X-ray emitting regions of different objects, like the atmosphere of cataclysmic variables, X-ray binaries, etc\\ldots. These codes are also used to model the absorption lines observed in the UV and X-ray range in quasars and AGN. Numerous new X-ray atomic data have been introduced in the codes, in order to obtain the best possible accuracy on the X-ray spectrum. But the ``escape probability approximation\" still lies at the center of the computation of the line intensities. This formalism, developed in the sixties and in the seventies, can be very useful to perform rapid approximate computations, but it does not lead to a correct estimation of the source function, especially in the case of strongly interlocked transitions, including continuum ones. Moreover it uses as a local quantity a global one, computed by an integration over the whole medium; this is a dramatic extrapolation for very inhomogeneous media like photoionized plasma. These aspects have been completely overlooked during the last twenty years, but they become now crucial in the context of X-ray emitting media. In a previous paper, Dumont et al. (2003, referred as D03) have shown that escape probability approximations, at least as they are used in the present codes, are unable to compute correct line intensities, within factors of ten, when the Thomson thickness of the medium is of the order of a few units, typical for the irradiated atmospheres of accretion discs in AGN and X-ray binaries. In the present paper, we extend this study to less optically thick media, and we show that the intensities of emission lines are also not accurately computed for parameters typical of the Warm Absorber of Seyfert 1 galaxies and of the region giving rise to the X-ray spectrum in Seyfert 2 galaxies \\footnote{One should note that modeling an absorption spectrum is much easier than an emission one. It requires only a correct computation of the thermal and ionization equilibrium, which give the fractional ion abundances, i.e. the populations of the ground levels, and thus the equivalent widths of the resonance lines, which can be compared to those deduced from the observations through a curve of growth analysis implying no line transfer. However it can happen (and this is indeed the case for the Warm Absorber), that emission lines are also produced by the absorbing medium, and one should thus worry about the line transfer.}. In the next section, we recall very briefly the essence of the problem, and we give a few examples in Section 3. ", "conclusions": "We have shown that the use of the escape probability approximation leads to large errors in the computed line fluxes and line ratios, in conditions which are typical of the X-ray emitting regions of Seyfert 2 nuclei, and of the Warm Absorber of Seyfert 1, i.e. for a Thomson thickness of the order of 0.001 or larger. This completes the previous paper (D03), where the same study was performed for thicker media, and where it was shown that the use of the escape approximation leads to errors by more than one order of magnitude on the line fluxes and line ratios. We find here that the errors are of the order of 30$\\%$ for a column density of 10$^{20}$ cm$^{-2}$, and can reach a factor five for a column density 10$^{23}$ cm$^{-2}$. We confirm that they are almost always in the direction of an overestimation of the most intense line intensities, especially of the L$\\alpha$ line of H-like ions, and of the resonance w term of He-like ions. We explain why such large errors occur for X-ray emitting media, and not for cooler media like the BLR. The comparison between the escape approximation and the transfer treatment (performed through the Accelerated Lambda Iteration method with our photoionization code Titan), is made in such a way that no other possible explanation of the discrepancies than the use of the escape probability approximation can be invoked. So one is led to conclude that unless a real transfer of the lines is introduced in the codes for modelling X-ray spectra, the results cannot have an accuracy better than that given by the approximation, even when the treatment of atomic physics is highly sophisticated. Whatever the discrepancy between the results of the escape probability and the transfer treatment, an uncertainty remains concerning the real intensity of the resonance lines. Here they were computed assuming complete redistribution within Doppler core, which mimics partial redistribution within a Voigt profile. PRD cannot be taken into account accurately with the escape probability formalism, but it can be done with the transfer treatment, provided the implementation in the code of another substantial time consuming procedure. We are presently studying such an improvement. \\bigskip \\noindent {\\bf APPENDIX: Equations used for the escape approximation in the optically thin case} We compute the escape probability towards the surface $P_{e}(\\tau_0)$ of a line as (cf. Collin-Souffrin et al. 1981): \\begin{equation} P_{e}(\\tau_0) ={\\rm max}(f1,f2), \\label{eq-esc-1} \\end{equation} for all subordinate and high resonance lines, and \\begin{equation} P_{e}(\\tau_0) =f1 \\label{eq-esc-2} \\end{equation} for the first resonance lines of H- and He-like species, with \\begin{eqnarray} f1 ={0.5\\over 1+2\\tau_0\\sqrt{\\pi ln(\\tau_0+1)}} && f2= {2\\over 3}\\sqrt{{a\\over\\sqrt{\\pi}\\tau_0}}, \\label{eq-esc-3} \\end{eqnarray} where $\\tau_0$ is the optical thickness at the line center between the emission point and the surface (we recall that it is taken along the photon path, so it multiplied by a factor $\\sqrt{3}$ to take into account the fact that the direction of the emitted photon is at random, and to be able to compare the escape approximation to the transfer treatment), and $a$ is the usual damping constant. $f1$ corresponds to the Doppler core, and $f2$ to the Lorentz wings of the Voigt profile. The total escape probability $P_{\\rm esc}$ is the sum of the escape probability towards the illuminated side, $\\beta_{\\rm ref}=P_{\\rm e}(\\tau_0)$, and towards the back side, $\\beta_{\\rm out}=P_{\\rm e}(T_0-\\tau_0)$, where $T_0$ is the total optical thickness of the slab depth at the line center. In the equations for the level populations, the net radiative rate from an excited level is replaced by $A_{\\rm ul}\\beta_{\\rm pop}$, with \\begin{equation} \\beta_{\\rm pop}={\\rm min}[1, P_{\\rm esc}(\\tau) \\times (1+{\\kappa_{\\rm c}\\over \\kappa_l}F({\\kappa_{\\rm c}\\over \\sqrt{\\kappa_{\\rm l}}}))], \\label{eq-betapop} \\end{equation} where $\\kappa_{\\rm c}$ and $\\kappa_{\\rm l}$ are respectively the absorption coefficient in the continuum and in the line, and $ F$ is the operator given by Hummer (1968) to account for destruction by continuum absorption in one line scattering: \\begin{equation} F(X)=\\int_{-\\infty}^\\infty {\\phi (x)\\over X+\\phi (x)} dx \\label{eq-FX} \\end{equation} where $\\phi (x)$ is the absorption line profile, $x=\\delta \\nu /\\delta \\nu _{\\rm D}$, and $\\delta \\nu _{\\rm D}$ is the Doppler width. A term $B_{\\rm lu}J_{\\nu}^{\\rm inc.att.}$ is added to the net radiative rate to take into account excitations by the attenuated incident radiation $J_{\\nu}^{\\rm inc.att.}=2\\beta_{\\rm ref}/(2\\pi)F_{\\nu}^{\\rm inc}$, according to the definition of the flux, $F_{\\nu}^{\\rm inc}$ being the incident flux at the frequency $\\nu$ (the corresponding deexcitation rate is negligible). The local cooling for each line is: \\begin{equation} \\Lambda_{\\rm line}=(n_{\\rm u}A_{\\rm ul} P_{\\rm esc}-n_{\\rm l}B_{\\rm lu}J_{\\nu}^{\\rm inc.att.}) \\ {h\\nu\\over n_{\\rm e}n_{\\rm H}} . \\label{eq-lambda-line} \\end{equation} The reflected flux in a line is computed as: \\begin{equation} F_{\\rm ref}= \\int{n_{\\rm u}A_{\\rm ul}h\\nu \\beta_{\\rm ref}\\ {\\rm exp}[-\\tau_{\\rm e}] dz}, \\label{eq-Fref} \\end{equation} where $\\tau_{\\rm e}$ is the effective optical thickness of the slab in the continuum at the line frequency, between the current point and the illuminated surface. In the outward emitted line flux, one must take into account the photons absorbed in the incident continuum: \\begin{eqnarray} F_{\\rm out}&=& \\int{(n_{\\rm u}A_{\\rm ul}h\\nu \\beta_{\\rm out} {\\rm exp}[-(T_{\\rm e}-\\tau_{\\rm e})]} \\\\ \\nonumber && \\ \\ \\ -n_{\\rm l}B_{\\rm lu}J_{\\nu}^{\\rm inc.att.}) dz, \\label{eq-Fout} \\end{eqnarray} where $T_{\\rm e}$ is the effective total optical thickness of the slab in the continuum at the line frequency. The ionization rate due to the lines, at the depth $z$, is equal to: \\begin{eqnarray} &&\\sqrt{3}\\kappa_{\\rm c} A_{\\rm ul}\\times \\\\ \\nonumber && \\left( \\int_0^z{(n_{\\rm u} \\beta_{\\rm out}{\\rm exp}[-\\tau_{\\rm e}+T_{\\rm e}]-n_{\\rm l}{\\rm B_{\\rm lu}\\over A_{\\rm ul}}J_{\\nu}^{\\rm inc.att.})} dZ \\right. \\\\ \\nonumber &&\\left. + \\int_H^z{n_{\\rm u} \\beta_{\\rm ref}{\\rm exp}[-T_{\\rm e}+\\tau_{\\rm e}]dZ}\\right). \\label{eq-ion} \\end{eqnarray} This expression is used also for the gains by photoionizations due to the lines. The escape probability is slightly different in the case of a highly ionized and/or moderately thick medium, as one must take into account the shift of line photons by comptonization. Thus $\\beta_{\\rm ref}$ is replaced by $\\beta'_{\\rm ref}= \\beta_{\\rm ref}+{1\\over 2} (1-P_{\\rm esc}(\\tau)) {\\sigma\\over \\kappa_l\\sqrt{\\pi}+\\kappa_c+\\sigma}$, and $\\beta_{\\rm out}$ by $\\beta'_{\\rm out}= \\beta_{\\rm out}+ {1\\over 2} (1-P_{\\rm esc}(\\tau)) {\\sigma\\over \\kappa_l\\sqrt{\\pi}+\\kappa_c+\\sigma}$, and $P_{\\rm e}$ is replaced by $P'_{\\rm e}=\\beta'_{\\rm ref}+\\beta'_{\\rm out}$." }, "0403/astro-ph0403523_arXiv.txt": { "abstract": "We present the long term X-ray light curves, detailed spectral and timing analyses of XTE J1908+094 using the Rossi X-ray Timing Explorer Proportional Counter Array observations covering two outbursts in 2002 and early 2003. At the onset of the first outburst, the source was found in a spectrally low/hard state lasting for $\\sim$40 days, followed by a three day long transition to the high/soft state. The source flux (in 2$-$10 keV) reached $\\sim$100 mCrab on 2002 April 6, then decayed rapidly. In power spectra, we detect strong band-limited noise and varying low-frequency quasi periodic oscillations that evolved from $\\sim$0.5 Hz to $\\sim$5 Hz during the initial low/hard state of the source. We find that the second outburst closely resembled the spectral evolution of the first. The X-ray transient's overall outburst characteristics lead us to classify XTE J1908+094 as a black-hole candidate. Here we also derive precise X-ray position of the source using Chandra observations which were performed during the decay phase of the first outburst and following the second outburst. ", "introduction": "The X-ray transient source \\xsrc was serendipitously discovered on 2002 February 19 during scheduled Rossi X-ray Timing Explorer (RXTE) Proportional Counter Array (PCA) observations of a Soft Gamma Repeater, SGR 1900+14 (Woods et al. 2002). Subsequent RXTE/PCA scanning observations of the region allowed the localization of the new source to RA: 19h08m50s, Dec: +09$^\\circ$22$\\arcmin$ 30$\\arcsec$ with an accuracy of 2$\\arcmin$. This placed the new source about 24$\\arcmin$ away from SGR 1900+14. Based on Very Large Array observations on 2002 March 21 and 22, a transient source was suggested as the radio counterpart candidate to \\xsrc (Rupen, Dhawan \\& Mioduszewski 2002a). Observations in the optical band on April 8 and 9 revealed no new sources near the radio position (Garnavich, Quinn \\& Callanan 2002), however Chaty, Mignani \\& Israel (2002) identified a near infrared counterpart to the new source and concluded that \\xsrc is in a low-mass X-ray binary system with a main sequence companion of spectral type later than K. Based on our preliminary analysis of the RXTE/PCA observations we concluded that \\xsrc is a new stellar mass black hole candidate (Woods et al. 2002). There are currently 18 dynamically confirmed and 20 candidate stellar mass black hole systems in our Galaxy (see McClintock \\& Remillard 2003 for a recent review). Most of them are characterized by occasional transient outbursts (X-ray novae; Chen, Shrader \\& Livio 1997) as a result of sudden increase in the mass accretion rate possibly triggered by instabilities in the accretion disk (Cannizzo 1993, Dubus et al. 2001). During outbursts, these systems generally undergo various changes in their spectral characteristics, usually in conjunction with changes in their timing behavior (see e.g., Homan et al. 2001). The most common BH spectral states are {\\it the low/hard state:} the spectrum is represented by a hard power law and usually accompanied by timing variability, and {\\it the thermal-dominant (high/soft) state:} a blackbody appears in the spectrum as the power law component gets steeper and timing features get weaker or completely disappear (Tanaka \\& Lewin 1995). Other spectral states characterized by more complicated spectral and timing properties are also observed (e.g., Homan et al. 2001). In this study, we present the results of our spectral and timing analysis of the RXTE pointed observations of \\xsrc covering two outburst episodes. Additionally, we report observations with the Chandra X-Ray Observatory taken under Director's Discretionary Time (DDT). We describe our observations in \\S 2, and we present detailed data analyses in \\S 3. In \\S 4 we discuss and compare the different states of the source to other black hole candidates. ", "conclusions": "In the course of its X-ray activity XTE J1908+094 proceeds through a series of X-ray states characteristic of black hole binaries, which we shall discuss using the terminology proposed by McClintock \\& Remillard (2003). The discovery outburst begins in the low/hard state, which lasted until about day 87. The source then enters an intermediate state during which the 2$-$5 keV flux peaks. This ends near day 90 when the source enters the thermal-dominant (high/soft) state, which persists until day 143, just before the secondary peak, where the source enters an intermediate state and then after day 149 returns to the low/hard state. The remaining observations (in 1.5$-$12 keV) up to the peak of the second outburst show the source in the low/hard state (see Figure \\ref{fig:rates_obss}). The one observation following the second outburst peak is consistent with the thermal-dominant state. In the low/hard state the energy spectra are dominated by a hard power-law component, and the power spectra by strong band-limited noise. During the low/hard state at the onset of the first outburst, the index of the power-law spectra began at $\\Gamma=1.4$, and then gradually softened to $\\Gamma=1.7$. The power spectra show band-limited noise with an rms amplitude which began near $r=30$\\% and gradually fell to $r=22$\\%. In addition there was a QPO with rms amplitude varying from 3\\% to 13\\%, which rose in frequency from 0.5 to 2.2 Hz. In the second interval of low/hard state ( days 149$-$295) the flux is lower, and the behavior of the power-law index is more complex. Due to the low flux, we could not make significant power-spectral measurements. Outburst onsets in the low/hard state have been seen in a number of X-ray novae. Brocksopp et al. (2002) tabulate 13 sources with outbursts that began in the low/hard state, five of which never left this state. Strong low-frequency QPO with rising frequencies are common in these low/hard state onsets, and have been seen for GRO J0422+32 \\cite{vanderHooft99}, GRO J1719-24 \\cite{vanderHooft96}, XTE J1550-564 (Finger et al. 1998; Cui et al. 1999), 4U 1630-472 \\cite{Dieters00}, XTE J1859+226 \\cite{Markwardt99}, and XTE J1118+480 \\cite{Wood00}, among others. \\begin{figure}[!b] \\centerline{\\includegraphics[scale=.50]{f16.eps}} \\vspace{0.0in} \\caption{\\baselineskip =0.5\\baselineskip The RXTE/PCA spectral fit for day 64 shown together with the BeppoSAX/PDA spectral fit for days 63$-$65 from in't Zand et al. (2002). \\label{fig:PDAspec}} \\end{figure} Hard-X-ray and gamma-ray observations have shown that in the low/hard state the power-law spectra break in the 100 keV range \\cite{Grove98}. In Figure \\ref{fig:PDAspec} we show our spectral fit for the PCA data on day 64 along with the spectral fit for BeppoSAX/PDA data from days 62$-$65 (MJD 52342$-$52345) \\cite{intZand02}. We notice a break near 50 keV. The flux in the 30-250 keV range is $3.2\\times 10^{-9}~{\\rm erg~cm}^{-2}~{\\rm s}^{-1}$, which surpasses the flux in the 2.5-25 keV range of $2.63\\times 10^{-9}~{\\rm erg~cm}^{-2}~{\\rm s}^{-1}$. The low/hard state is also associated with radio emission. Flat spectrum radio emission, associated with compact jets, is consistently observed during the low/hard state of X-ray novae \\cite{Fender03}. Indeed, during the onset of the first XTE J1908+094 outburst, Very Large Array observations (on days 74$-$75) led to the discovery of a radio counterpart to XTE J1908+094, with a flux of 0.85 mJy at 8.6 GHz \\cite{Rupen02a}. This was detected in additional observations until day 127 \\cite{Rupen02b}. \\begin{figure}[!b] \\centerline{\\includegraphics[scale=.50]{f17.eps}} \\vspace{0.0in} \\caption{\\baselineskip =0.5\\baselineskip Flux, Power-law index, and variability evolution during the intermediate state. The top panel shows the 2.5-25 keV flux for the power-law (filled triangles), the disk black-body component (filled circles), and their sum (dashed). The middle panel shows the power-law index. The lower panel shows the variability amplitude.\\label{fig:IMstate}} \\end{figure} Figure \\ref{fig:IMstate} shows the X-ray flux, power-law index and variability amplitude evolution during the transition interval between the low/hard and thermal-dominant state. This intermediate state begins near day 67, when the power-law flux begins to drop, the power-law component begins to rapidly soften, and the disk black-body flux begins to rise, and the flux variability begins to fall. The rise of the disk black-body flux occurs in four days, but the fall of the power-law flux takes 15 days to complete. BeppoSAX/MECS observations covering day 66.4 $-$67.8 (MJD 52366.4$-$52367.8) show the onset of this transition \\cite{intZand02}. In this transition, the total flux in 2.5$-$25 keV band consistently falls. However, a significant fraction of the disk black-body flux is below this energy range. From our spectral fits we find that the total integrated flux of this thermal component rises to $\\sim 7\\times 10^{-9}~{\\rm erg~cm}^{-2}~{\\rm s}^{-1}$, implying that the bolometric flux may be constant or rising. After day 90 the source is in the thermal-dominant state, with thermal disk flux dominating the spectrum, and low variability. The disk black-body normalizations average about 40, which is consistent with the inner disk being at the radius of the inner most stable circular orbit if \\begin{equation} (M/M_\\odot)D_{10kpc}^{-1}\\cos^{\\onehalf}\\theta \\approx 4.5 \\end{equation} where $M$ is the black-hole mass, $D_{10kpc}$ the source distance in units of 10 kpc, and $\\theta$ the disk inclination to the line of sight. Note here that for any value of disk inclination angle, the mass of the central object is in the range of a black hole, if the source distance is of the order of 10 kpc. During this thermal-dominant state, the flux steadily falls. This mainly occurs by the temperature decreasing. The thermal-dominant state ends on day 149 when a transition begins back to the low/hard state. In the transition from the low/hard state to the thermal-dominant state starting day 87), the disk black-body normalization begins near 15 and rises, implying an increasing inner disk radius. The opposite occurs on the transition back to the low/hard state. This is counter to the expectation that the inner-disk radius is large during the low/hard state, and near the inner most stable orbit in the thermal-dominant state (e.g., Esin et al. 1997). This rise and fall may be due to systematic problems with our spectral fits: we detect only the high-energy tail of the thermal spectrum. In fits where the column density, disk temperature and flux are all free to vary, these parameters are, therefore, highly coupled. By fixing the column density to the value found with the BeppoSAX data, we have reduced this coupling, but could be biasing the solution. There is a strong correlation between the flux associated with the broad line feature and that of the power law component during the early stages of the first outburst episode. One possible interpretation of this feature is that it is the fluorescent Fe K$\\alpha$ emission produced by the reprocessing of the hard X-ray photons by cooler material close to the central object. The line centroid energies were somewhat lower than what is expected for neutral iron (6.4 keV). This may suggest that what we observe is primarily the red wing of Doppler shifted neutral Fe K$\\alpha$ in a Keplerian accretion. This was seen also in 4U 1630$-$47 (Cui, Chen \\& Zhang 1999) and in XTE J1748$-$288 (Miller et al. 2001). Esin et al. (1997) have presented a model to explain the states of X-ray novae. In the quiescent and low/hard state a thin accretion disk is present but truncated at a large inner radius. Within this radius there is an Advection Dominated Accretion Flow (ADAF) which is a hot and radiatively inefficient flow where most of the thermal energy generated is advected onto the black hole rather then being radiated. Above the accretion disk is a hot corona, which is a continuation of the advection dominated flow, which produces a power-law spectral component in the low/hard state by Comptonization. This model does not incorporate the jets which are responsible for the radio emission now known to be associated with the low/hard state. Markoff, Falcke \\& Fender (2001) have proposed that these jets also produce the power-law component via synchrotron radiation. In their model a standard accretion disk transitions at an inner radius of $\\sim 10^3$ km to a hot ADAF-like flow which feeds the jet. The hard X-rays are synchrotron radiation produced in a shock acceleration region some $10^3$ km above the disk plane. The radio emission is from beyond this region. While providing successful fits of a multi-wavelength spectrum, neither of these models yet consider dynamical changes in the flow or attempt to explain the power-spectra seen in the different states. The high amplitude variability seen in the low-hard state requires changes in emissivity that are spatially coherent over most of the emission region. It is tempting to associate the QPO's seen in the low/hard state with the Keplerian frequency at the inner edge of the thin accretion disk. Yet while this is plausible in the low-hard state, the QPO in XTE J1908+094 persists into the intermediate state, where the inner disk radius inferred from spectral fits imply frequencies much larger than those observed." }, "0403/astro-ph0403715_arXiv.txt": { "abstract": "We present spectroscopic confirmation of the Cl 1604 supercluster at $z \\sim 0.9$. Originally detected as two individual clusters, Cl 1604+4304 at $z = 0.90$ and Cl 1604+4321 at $z = 0.92$, which are closely separated in both redshift and sky position, subsequent imaging revealed a complex of red galaxies bridging the two clusters, suggesting that the region contained a large scale structure. We have carried out extensive multi-object spectroscopy, which, combined with previous measurements, provides $\\sim600$ redshifts in this area, including 230 confirmed supercluster members. We detect two additional clusters that are part of this structure, Cl 1604+4314 at $z = 0.87$ and Cl 1604+4316 at $z = 0.94$. All four have properties typical of local clusters, with line-of-sight velocity dispersions between 489 and 962~{\\kms}. The structure is significantly extended in redshift space, which, if interpreted as a true elongation in real space, implies a depth of $93~h_{70}^{-1}~{\\rm Mpc}$. We examine the spatial and redshift distribution of the supercluster members. ", "introduction": "The Cl 1604 supercluster was initially detected as two separate clusters, Cl 1604+4304 at $z=0.90$ and Cl 1604+4321 at $z=0.92$, in the plate-based survey of \\citet{gho86}. Deeper imaging and multi-object spectroscopy taken with the Low Resolution Imaging Spectrograph \\citep[LRIS;][]{oke95} at the Keck telescopes by \\citet[hereafter O98]{oke98} yielded redshifts and preliminary velocity dispersions for each cluster \\citep[][hereafter P98,P01]{pos98,pos01}. Motivated by the close separation of Cl 1604+4304 and Cl 1604+4321 in both radial velocity (4300~{\\kms}) and position on the sky ($17'$), \\citet{lub00} performed deep multi-band imaging with the Palomar 5-m telescope covering the area between the two clusters. The imaging revealed an overdensity of red galaxies whose colors were consistent with early-type galaxies at $z \\sim 0.9$, suggesting that the clusters were part of a high-redshift supercluster. To verify this conclusion, map the supercluster structure, and obtain a large sample of cluster galaxies with measured spectroscopic properties, we are conducting an extensive spectroscopic survey spanning the region between Cl 1604+4304 and Cl 1604+4321. In this Letter, we present spectroscopic confirmation of this large scale structure and a discussion of its properties based on the currently known supercluster members. We assume a $\\Lambda$CDM cosmology with $\\Omega_m=0.3, \\Lambda=0.7,$ and ${\\rm H_0}=70~h_{70}~{\\rm km~s^{-1}~Mpc^{-1}}$. ", "conclusions": "We have obtained the largest spectroscopic sample to date for a supercluster at $z\\sim0.9$. The initial results suggest that the four component clusters are part of a single, filamentary structure, with the individual clusters being massive (Abell Richness class 0--2). As seen in Figure~\\ref{gal_posns}, we may not yet have reached the outer limits of this structure, and future wide-area imaging and additional spectroscopy will be required to measure its full extent. Already our dataset is comparable in size to studies of local superclusters (see \\S3.2). This structure is by far the best studied of the very few known superclusters at $z>0.3$ \\citep{con96,ros99,gav04}. The spectroscopic data, especially from DEIMOS, are of sufficient resolution and quality to measure equivalent widths for both emission and absorption features. A large fraction of the galaxies in the cluster show evidence for star-formation (most notably strong [OII] emission). Examining the correlation of such features with galaxy location will yield insights into the dynamics and history of the supercluster system and shed light on the role of cluster-scale physical processes (e.g., shocks and ram-pressure stripping) in galaxy star formation. Proposed HST imaging and Chandra X-ray mapping of this structure will allow us to compare galaxy morphological and spectroscopic properties and correlate optical substructure with the location and temperature of the hot gas component. These data, combined with numerical simulations, may be able to predict future mergers and perhaps even the end-state of this system. This survey will provide a view of a young supercluster rivaling the detailed studies of local structures." }, "0403/astro-ph0403465_arXiv.txt": { "abstract": "Hard X-ray surveys have uncovered a large population of heavily obscured AGN. They also reveal a population of quasars with moderate obscuration at both visible and X-ray wavelengths. We use \\chandra selected samples of quasars from the ELAIS Deep X-ray Survey (EDXS) and the \\chandra Deep Field-North to investigate the obscuration towards the nuclei of moderately obscured AGN. We find an inverse correlation between the optical to X-ray flux ratio and the X-ray hardness ratio which can be interpreted as due to obscuration at visible and X-ray wavelengths. We present detailed optical and near-infrared data for a sample of optically-faint ($R>23$) quasars from the EDXS. These are used to constrain the amount of rest-frame UV/optical reddening towards these quasars. It is found that optically-faint quasars are mostly faint due to obscuration, not because they are intrinsically weak. After correcting for reddening, the optical magnitudes of most of these quasars are similar to the brighter quasars at these X-ray fluxes. Combining with gas column densities inferred from the X-ray observations we consider the gas-to-dust ratios of the obscuring matter. We find that the quasars generally have higher gas-to-dust absorption than that seen in the Milky Way -- similar to what has been found for nearby Seyfert galaxies. We consider the possible existence of a large population of X-ray sources which have optical properties of Type 1 (unobscured) quasars, but X-ray properties of Type 2 (obscured) quasars. However, we find that such sources only contribute about 6\\% of the 0.5-8 keV X-ray background. Finally we show that the observed distribution of optical-to-X-ray flux ratios of quasars at $z>1$ is skewed to low values compared to the intrinsic distribution due to the fact that the observed-frame $R$-band light is emitted in the UV and is more easily obscured than hard X-rays. ", "introduction": "The hard spectral shape of the X-ray background led to the idea that a large population of obscured active galactic nuclei (AGN) exist which fail to show up in optically-selected quasar surveys (Comastri et al. 1995). Deep hard X-ray surveys with \\chandra and \\xmm have now revealed the sources responsible for the 0.5-10 keV X-ray background (e.g. Hornschemeier et al. 2001; Tozzi et al. 2001; Hasinger et al. 2001; Manners et al. 2003). As expected, the majority of the optical counterparts to hard X-ray sources are galaxies containing optically-obscured AGN. Seyfert galaxies at low redshift often show complex absorption structures consisting of both cold (neutral) and warm (partially ionized) absorbers (Mushotzky, Done \\& Pounds 1993). A fraction of Seyferts appear to be Compton thick and their observed X-ray emission is reflection dominated. Due to the low signal-to-noise of the X-ray spectra of the sources responsible for the X-ray background, there is only limited knowledge of their absorption properties. It is unknown whether the orientation-based obscuration scheme which works well for low-redshift Seyferts can also be applied at higher redshifts. There are still only a handful of Type 2 narrow-line quasars found in deep X-ray surveys (Almaini et al. 1995; Norman et al. 2002; Stern et al. 2002; Crawford et al. 2002; Mainieri et al. 2002; Szokoly et al. 2004), seemingly at odds with simple unification schemes (however there are still many faint objects without redshifts in these surveys which could be Type 2 quasars). The relative lack of luminous Type 2 quasars and the dominance of obscured AGN at lower luminosities (Ueda et al. 2003; Hasinger et al. 2003) suggests that the fraction of obscured objects is a strong function of luminosity, as had previously been found for low-radio-frequency selected AGN (Simpson, Rawlings \\& Lacy 1999; Willott et al. 2000). The fraction of Compton thick AGN in the X-ray background sources is also quite uncertain (e.g. Fabian, Wilman \\& Crawford 2002). The dust properties of the high redshift X-ray absorbing material have not yet been well studied. There are certainly objects at low redshift with dust absorption quite different from that expected from the observed X-ray absorption based on a Galactic gas-to-dust ratio (Simpson 1998; Maiolini et al. 2001a) and a few cases of such discrepancies at higher redshifts have been reported (Akiyama et al. 2000; Risaliti et al. 2001; Willott et al. 2003; Watanabe et al. 2004). In this paper we discuss the dust and X-ray absorption present in hard X-ray sources contained within the ELAIS Deep X-ray Survey and the implications for understanding the sources responsible for the X-ray background. In Sec.\\,2 we discuss the observed correlations between optical and X-ray properties of X-ray selected quasars and what these suggest about obscuration. Sec.\\,3 presents near-infrared and optical data for optically-faint X-ray selected quasars. In Sec.\\,4 we fit model quasar spectra to the observations to constrain the reddening towards these quasars. In Sec.\\,5 we compare the obscuration in the UV/optical with that in X-rays to determine the gas-to-dust ratio of the obscuring material and compare this with values in the Milky Way, the Small Magellanic Cloud and low-redshift AGN. In Sec.\\,6 we discuss the effect that obscuration plays in altering the intrinsic optical-to-X-ray flux ratios of quasars to those observed. We assume throughout that $H_0=70~ {\\rm km~s^{-1}Mpc^{-1}}$, $\\Omega_{\\mathrm M}=0.3$ and $\\Omega_\\Lambda=0.7$. ", "conclusions": "Deep X-ray surveys are capable of uncovering the AGN responsible for most of the supermassive black hole accretion history of the universe. Combined X-ray and optical/near-infrared observations of lightly reddened quasars from such X-ray surveys are a powerful probe of the physical conditions of the obscuring material. We have used optical and near-infrared photometry and spectroscopy to constrain the amount of reddening in the light from optically-faint quasars. Our main conclusions are: \\begin{itemize} \\item{Optically-faint ($R>23$) quasars at $f_{0.5-8}>2 \\times 10^{-15}$\\,erg\\,cm$^{-2}$\\,s$^{-1}$ are mostly faint due to obscuration. De-reddening the observed $R$-band fluxes of these quasars gives them optical magnitudes similar to other quasars with these X-ray fluxes ($201$ is skewed to low values compared to the intrinsic distribution due to the fact that the observed-frame $R$-band light is emitted in the UV and is more easily obscured than the hard X-rays sampled by {\\it Chandra}.} \\end{itemize}" }, "0403/astro-ph0403186_arXiv.txt": { "abstract": "{We use Kuiper's test to detect periodicities in X-ray and gamma-ray observations. Like Rayleigh's test, it uses the individual photon arrival times, and is therefore well suited to the analysis of faint sources. Our method makes it possible to take into account the discontinuities in the observation, and to completely get rid of the contamination that results from them. This makes it particularly adapted to the search of periods long compared to the total observation duration. We propose a semi-analytical approach to determine the effective number of trial frequencies when searching for unknown periods over a frequency range. This approach can be easily adapted to other tests. We show that, using Kuiper's test, we can recover periods in frequency domains where other tests are completely confused by contamination. We finally search the entire ROSAT Position-Sensitive Proportional Counter (PSPC) archive for long periods, and find 28 new periodic-source candidates. ", "introduction": "\\label{sec:intro} Important efforts have been devoted to the search of periodic signals throughout the electromagnetic spectrum. Because of the idiosyncrasies of astrophysical observations, different methods must be used depending on the type of object and the wavelength range. Four test families seem to dominate the period-detection ``market''. The calculation of the Fourier power spectrum density \\citep[e.g.,][]{PresEtal-1993-NumRec} using a fast Fourier transform (FFT) is adapted to evenly spaced (or evenly binned) observations. The Lomb-Scargle periodogram \\citep{Lomb-1976-LeaSqu,Scar-1982-StuAstII,HornBali-1986-PrePer}, a discrete Fourier transform method, can be used for unevenly-spaced flux measurements. Epoch folding (EF) \\citep[e.g., ][]{LeahEtal-1983-SeaPul} can be used in the same conditions or for individual photons, but requires a binning according to the phase. Rayleigh's test \\citep[e.g.,][]{GibsEtal-1982-TraEmi,Fish-1993-StaAna} is particularly adapted for the analysis of individual photons. Observations in the X- and gamma-rays usually have two important characteristics. First, independent, time-tagged photons are collected. A method requiring binning is therefore far from ideal, as it results in a loss of information. Furthermore, binning is prohibited for sources detected with very few photons; for EF for instance, the required assumption of Gaussian distribution in each bin is not satisfied in this case. Moreover, the necessary assumptions on the number and sizes of the bins lower the performance of the test \\citep{Schw-1999-OptPer}. Secondly, space observations are often interrupted by ``bad time'' periods, where no data are received. Fourier-based methods and Rayleigh tests are seriously affected by this problem. In practice, it means that only periods short compared to the durations of uninterrupted observation can be investigated. In this paper we present in detail Kuiper's test \\citep{Kuip-1960-TesCon}. This test has been applied to the distribution of solar flares \\citep{JetsEtal-1997-LonDis}, and to the search for periodicities in Earth impacts \\citep{Jets-1997-HumSta,JetsPelt-2000-SpuPer}, but its unique suitability to X-ray and gamma-ray observations has been overlooked. Similarly to Rayleigh's test, it uses discrete events, and can be applied to very faint sources without any {\\em a priori} assumption. Similarly to EF, it takes into account non-uniform coverage of the phase domain, and can therefore be used when searching for periods long compared to the total observation duration\\footnote{In this paper, ``total duration'' means the time interval between the start and the end of an observation, including possible gaps.}. We study in detail the properties of Kuiper's test for period detection, and particularly its significance level. We concentrate on two important issues: the treatment of discontinuous observations, and the determination of the effective number of trial frequencies when searching for unknown periods. We finally apply the algorithm to the entire archive of the ROSAT Position-Sensitive Proportional Counter (PSPC) archive. ", "conclusions": "\\label{sec:conc} Kuiper's test shows very interesting properties for the search of long-period periodic objects. Its ability to cope very naturally, without any hidden assumption, with complex GTIs is unique. Compared to Rayleigh's, Kuiper's test performs better for narrow-peaked light curves. Kuiper's test is quite sensitive to both subharmonics and harmonics of the fundamental frequency, but usually identifies the fundamental correctly. Kuiper's test is particularly adapted to X-ray missions, like XMM-Newton and Chandra, high-energy gamma-ray satellites like GLAST, and for Cherenkov telescopes. The semi-analytical method we propose here to correct the false-positive probability in case of a search over a range of frequencies should be quite useful in practice, not only for Kuiper's test, but also for other tests, as its principle can be easily adapted. It has the advantage of simplicity, and of being based on sound probability principles. On the 28 candidate periodic sources, 6 could be cross-checked using other ROSAT PSPC observations. Good or partial confirmation of the existence of periodicities is found in 3 of these objects, and there is total absence of confirmation in 3 objects. This does not necessarily imply a ``confirmation of absence''. It must be reminded that X-ray sources are quite often strongly variable, and that a periodic signal may remain undetected in some observations, even though the observing conditions seem adequate. For instance, \\citet{IsraEtal-2000-BepCha} report the detection of a periodic signal in the X-ray pulsar \\object{2E~0053.2-7242} in only one out of nine ROSAT PSPC observations, the source having dimmed by a factor $>\\!\\!6$ between the different observations. The possibility that extrinsic contamination, or statistical flukes explain some, or even most, of the candidate periods must be considered seriously. Firm identification of the candidates as periodic sources will be contingent upon the detection of the periods in independent data sets. The building up of important X-ray archives from XMM-Newton and Chandra makes it quite probable that new observations will be available for a fair number of these sources in the near future. A C library implementing the algorithms discussed in this paper is available from the author." }, "0403/astro-ph0403275.txt": { "abstract": "We combine the recently published CIZA galaxy cluster catalogue with the XBACs cluster sample to produce the first all-sky catalogue of X-ray clusters in order to examine the origins of the Local Group's peculiar velocity without the use of reconstruction methods to fill the traditional Zone of Avoidance. The advantages of this approach are (i) X-ray emitting clusters tend to trace the deepest potential wells and therefore have the greatest effect on the dynamics of the Local Group and (ii) our all-sky sample provides data for nearly a quarter of the sky that is largely incomplete in optical cluster catalogues. We find that the direction of the Local Group's peculiar velocity is well aligned with the CMB as early as the Great Attractor region 40 $h^{-1}$ Mpc away, but that the amplitude of its dipole motion is largely set between 140 and 160 $h^{-1}$ Mpc. Unlike previous studies using galaxy samples, we find that without Virgo included, roughly $\\sim70\\%$ of our dipole signal comes from mass concentrations at large distances ($>60$ $h^{-1}$ Mpc) and does not flatten, indicating isotropy in the cluster distribution, until at least 160 $h^{-1}$ Mpc. We also present a detailed discussion of our dipole profile, linking observed features to the structures and superclusters that produce them. We find that most of the dipole signal can be attributed to the Shapley supercluster centered at about 150 $h^{-1}$ Mpc and a handful of very massive individual clusters, some of which are newly discovered and lie well in the Zone of Avoidance. ", "introduction": "Since the detection of a dipole anisotropy in the Cosmic Microwave Background (CMB, Kogut 1993), many question have been raised regarding the origin of the Local Group (LG) motion which is thought to give rise to such a signal. Specifically at question have been the nature of the cosmic objects or structures most directly responsible for inducing the LG's peculiar velocity and, furthermore, the distance out to which inhomogeneities in the distribution of these objects continue to affect the LG's dynamics. Interest in these questions is motivated by the cosmological implications they carry; for example, if the motion of the Milky Way (MW) is induced entirely nearby, then the current density of the universe would need to be considerable in order for nearby matter concentrations to adequately accelerate the MW over such relatively small scales. On the other hand, if the LG's peculiar velocity is induced by more distant structures, then we know that anisotropies in the large-scale matter distribution must exist to at least those structures and that the universe becomes isotropic only at larger distances. To answer these questions, much effort has been spent examining the peculiar velocity that would be induced onto the LG by the distribution of various mass tracers, such as galaxies and clusters of galaxies, and comparing this motion with that inferred from the CMB. Traditionally such analyses have made use of the linear theory of gravitational instability, which dictates that the peculiar velocity of a reference frame can be related to the gravitational acceleration induced by the mass distribution surrounding it via (Peebles 1976, see our Appendix A) \\begin{equation} \\textit{\\textbf{v}}_{p} \\hspace{.1in} = \\hspace{.1in} \\frac{H_{o}\\beta}{4\\pi \\bar{n}} \\int \\frac{n(r)}{r^{2}} \\textit{\\textbf{\\^{r}}}\\ dr \\end{equation} \\noindent where $\\beta = \\Omega^{0.6}_{0}/b$ and $b$ is the biasing factor relating the mass-tracers to the underlying mass distribution they represent, and $\\bar{n}$ is the average mass-tracer number density. In other words, Equation 1 tells us that the dipole moment of a mass-tracer distribution can be directly related to the peculiar velocity that sample would induce on the LG. Within this framework, the comparison of the LG's peculiar velocity as inferred from the CMB dipole anisotropy to that produced by a mass-tracer distribution can shed light on the role of the sample on producing the LG's motion, provide an estimate of the depth at which inhomogeneities in the distribution of the sample affect the LG's dynamics (i.e. the convergence depth, $R_{conv}$), and place constraints on how the sample traces the underlying matter distribution in the form of the biasing parameter $\\beta$. Performing this type of analysis is a nontrivial matter for numerous reasons: (i) for linear perturbation theory to be applicable, the dipole moment of the sample under study must be relatively well aligned with the observed peculiar velocity of the LG; (ii) the characteristic depth of the sample must be larger than the depth at which anisotropies in the sample can influence the LG's dynamics ($R_{*} > R_{conv}$); (iii) the number of mass-tracers must be sufficient to accurately sample the underlying density field and avoid the introduction of shot-noise errors. Wary of these concerns, this type of dipole analysis has been extensively applied to the LG. While studies have used samples ranging from optical galaxies (Lahav, Rowan-Robinson \\& Lynden-Bell 1988; Lynden-Bell, Lahav \\& Burstein 1989; Shaya et al. 1992; Hudson 1993) to X-ray selected AGN (Miyaji \\& Boldt 1990), much of the early dipole work focused on IRAS galaxies due to their considerable sky coverage (Meiksin \\& Davis 1986; Yahil, Walker \\& Rowan-Robinson 1986; Strauss \\& Davis 1988; Yahil 1988; Rowan-Robinson et al. 1990, 1991; Strauss et al. 1992; Plionis, Coles \\& Catelan 1993; Basilakos \\& Plionis 1998). Despite the wide employment of such galaxy samples, they are all plagued by steeply declining selection functions with distance, leading to a significant incompleteness at large distances and characteristic depths of 80-100 $h^{-1}$ Mpc. Analyses of galaxy samples have produced dipoles that are generally well aligned with the CMB dipole direction, but many differ regarding their implied convergence depth. Strauss et al. (1992) and Hudson (1993) found an $R_{conv}$ of roughly $\\sim$$50 h^{-1}$ Mpc, implying that all of the peculiar acceleration of the LG is induced relatively nearby. On the other hand, Plionis (1988), Plionis, Coles \\& Catelan (1993), Basilakos \\& Plionis (1998), Branchini et al. (1999), Schmoldt et al. (1999), and Plionis et al. (2000) found evidence for contributions ranging from 100 to 150 $h^{-1}$ Mpc. The credibility of many of these early results must be questioned since the convergence depth of each sample shows a strong dependence on the sample's characteristic depth, a feature to be expected if anisotropies beyond that characteristic depth continued to contribute to the motion of the LG. This implies that galaxy samples alone do not probe the mass fluctuation field deep enough to account for all the anisotropies that affect the LG's dynamics. The use of clusters of galaxies overcomes some of the problems faced by galaxy samples since clusters are luminous enough for samples to be volume-limited out to larger distances. Although clusters only sparsely sample the underlying density field, which introduces shot-noise errors, they trace the peaks of the density fluctuation field, which have the greatest effect on the dipole amplitude and direction. Using the Abell/ACO cluster catalogue (Abell 1958, Abell, Corwin \\& Olowin 1989, hereafter ACO), which has a characteristic depth of over 240 $h^{-1}$ Mpc, evidence has been found for contributions to the LG dipole from depths of $\\sim$160 $h^{-1}$ Mpc (Scaramella, Vettolani \\& Zamorani 1991,1994; Plionis \\& Valdarnini 1991; Brunozzi et al. 1995; Branchini \\& Plionis 1996). These results were confirmed by Plionis \\& Kolokotronis (1998) using the X-ray Brightest Abell-type Cluster catalogue (XBACs, Ebeling et al. 1996), which is optically selected, but X-ray confirmed, thus eliminating projection effects which may have enhanced the dipole amplitude obtained with Abell/ACO (Sutherland 1988). In the context of the dipole analysis, the primary limitation of the XBACs and Abell/ACO catalogues are their incompleteness at low Galactic latitudes. This is because traditional optical cluster searches suffer from severe extinction and stellar obscuration in the direction of the Milky Way (MW), leading to catalogues with poor coverage in a $40^{\\circ}$ wide strip centered on the plane of the Galaxy, known as the Zone of Avoidance (ZOA). This is particularly troubling since Shaya suggested as early as 1984 that large-scale structures obscured by the ZOA could have a significant effect on the peculiar motion of the LG. More recently, several studies have found renewed evidence for a significant bulk motion toward a vertex behind the plane of the MW (Riess et al 1997; Hudson et al. 1999), rekindling the idea of a Great Attractor (Lynden-Bell et al. 1988). A variety of techniques have been used to reconstruct the ZOA, ranging from a uniform filling (Strauss \\& Davis 1988; Lahav 1987) to a spherical-harmonics approach which extends structures above and below the plane into the ZOA (Plionis \\& Valdarnini 1991, cf. Brunozzi et al. 1995). The value of these reconstruction techniques is, however, limited if the MW does indeed obscure dynamically significant regions, as has been suggested. With the recent compilation of the X-ray selected CIZA cluster catalogue (named for Clusters in the Zone of Avoidance; Ebeling, Mullis \\& Tully 2002), it has become possible to add real cluster data to the region behind the MW. In this study we combine the CIZA and XBACs samples, with appropriate weightings, to produce the first all-sky catalogue of X-ray luminous clusters and provide a dipole analysis of the LG peculiar velocity without the use of reconstruction methods to fill the ZOA. We proceed in the following manner: in section 2 we give an overview of the XBACs and CIZA samples, with a thorough look at the systematic effects in each. Section 3 describes the details of the dipole analysis and our results are put forward in section 4. Finally we summarize our primary conclusions in section 5. Throughout this paper we assume an Einstein-de Sitter universe with $q_{0}=0.5$ and $H_{0}=100$ $h$ km s$^{-1}$Mpc$^{-1}$, so that our results are directly comparable to those of previous studies. \\begin{figure}[t] \\hspace*{-0.15in} \\centerline{\\psfig{file=f1.eps,width=3.25in,angle=0}} \\caption{Aitoff projection of the XBACs and CIZA cluster catalogues in Galactic Coordinates. The dashed lines represent the traditional Zone of Avoidance ($|b|<20^{\\circ}$).} \\end{figure} ", "conclusions": "One of the limitations to the use of clusters of galaxies to identify the structures inducing the LG's peculiar motion has been the incompleteness of cluster catalogues at low Galactic latitudes. In this study we have combined the recently published CIZA catalogue with the XBACs cluster sample to produce the first all-sky catalogue of X-ray bright clusters in order to analyze the origin of the LG peculiar velocity without the need for reconstruction methods to fill the traditional ZOA. We find that the cluster distribution becomes isotropic with respect to the LG at a distance of 160 $h^{-1}$ Mpc and that with Virgo excluded the asymptotic value of the dipole amplitude is largely set by a coherent signal near $\\sim150 h^{-1}$ Mpc. While this agrees with previous results, our finding that $\\sim70\\%$ of the dipole amplitude is set at distances larger than $60 h^{-1}$ Mpc differs from results obtained using galaxy and optically selected cluster samples. By examining the dipole profile on a cluster-by-cluster basis, we conclude that the Shapley concentration is the single supercluster most responsible for producing the dipole signal between 140 and 160 $h^{-1}$ Mpc. We also find that the cluster dipole is fairly well aligned with the direction of the CMB dipole at relatively shallow distances, and that the competing effects of Virgo cluster and the PC concentration essentially counteract each other's effects on the final dipole pointing. These facts, coupled with the significant contributions to the LG peculiar acceleration from larger distances, reaffirm the bootstrap theory which suggests that the aligned mass concentrations of the GA and Shapley regions largely set the direction of the LG's peculiar motion throughout our study volume. Furthermore our analysis has identified four dynamically interesting CIZA clusters: C1324 in the GA region, C1410 in the Shapley supercluster and the two clusters C1652 and C1638 (Triangulum Australis) which lie behind the GA region at roughly the same distance as Shapley but well in the ZOA. Lastly, using the LG's peculiar velocity as measured from the CMB anisotropy (corrected for Virgo-centric infall) and the amplitude of the cluster dipole without Virgo included, we find the average value of the $\\beta$ parameter to be $\\beta = 0.24 \\pm 0.02$, in agreement with previous values determined from the XBACs catalogue alone. %On a final cautionary note, we remind the reader that including Virgo in our analysis highlighted the fact that our sample does note trace the mass distribution in the local neighborhood very well. %On a final cautionary note, we remind the reader that since our sample does not include the most nearby clusters such as Virgo, it does not constrain the dipole zero-point very well. %On a final cautionary note, we remind the reader that our sample does not include some of the nearest clusters such as Virgo, which causes our dipole zero-point to be ill-constrained. On a final cautionary note, we remind the reader that the sparseness of X-ray bright clusters causes them to be noisy tracers of the mass distribution when very small volumes are considered, which inturn causes the dipole zero-point to be ill-constrained. Therefore one should bear in mind that our conclusions regarding the relative contributions of different distance scales to the final dipole amplitude are valid only in the range we sample well: $40$ to $240 h^{-1}$ Mpc. That being said, we can think of two methods by which our results can be extended to shallower distances: (i) since truly X-ray selected cluster catalogues such as the BCS have been shown to better sample the nearby cluster distribution, the use of an all-sky, X-ray selected cluster sample would allow a dipole analysis to be performed without the need for weights to compensate for the residual incompleteness in the XBACs sample. The construction of such a sample will be possible in the near future when the REFLEX catalogue becomes available; its combination with the BCS and CIZA samples will produce the first entirely X-ray selected, all-sky data set. (ii) Nearby galaxy catalogues can be used to establish a zero-point for the X-ray cluster dipole, from which the relative contributions of different distance scales can truly be determined. This galaxy dipole normalization has been applied to optical cluster catalogues (Scaramella et al. 1994), but has yet to be implemented on a X-ray selected cluster sample." }, "0403/astro-ph0403379_arXiv.txt": { "abstract": "By integrating the relativistic hydrodynamic equations introduced by Taub we have determined the exact EQuiTemporal Surfaces (EQTSs) for the Gamma-Ray Burst (GRB) afterglows. These surfaces are compared and contrasted to the ones obtained, using approximate methods, by \\citet{pm98c,s98,gps99}. ", "introduction": "The recent explanation of the observed luminosity in X- and $\\gamma$-ray energy bands in Gamma-Ray Bursts (GRBs), as well as the comprehension of their spectral properties, depends in an essential way on the determination of the EQuiTemporal Surfaces (EQTSs) in the afterglow era \\citep{rbcfx02_letter,Spectr1}. Here we compare and contrasts the exact determination of the EQTSs with the approximate expressions presented by \\citet{pm98c,s98,gps99}. A great deal of consensus has been reached concerning three basic issues in the description of GRB afterglows: \\\\ {\\bf a)} Their origin is generally traced back to the interaction of an ultrarelativistic baryonic matter pulse with the InterStellar Medium (ISM). It is also agreed that the relativistic hydrodynamic equations introduced by Abe \\citet{Taub} are the correct theoretical framework to describe such an interaction.\\\\ {\\bf b)} The general definition of the EQTSs is also agreed upon by everyone.\\\\ {\\bf c)} The necessity of determining the boundary conditions of the baryonic matter pulse in the early phases of the afterglow era is also generally recognized. We illustrate in the following sections the equations needed for the description of these three basic issues and also point out a major difference between our approach and the ones in the current literature concerning the solutions of the Taub equations. We identify in this difference the reason for the results on the EQTSs presented in Fig.~\\ref{eqts_comp_ad} and Fig.~\\ref{eqts_comp_rad}. Our treatment assumes the exact integration of the Taub equations. We recall that all the GRB observable quantities depend essentially on the EQTSs. Due care is therefore needed in their correct determination. ", "conclusions": "The consequences of using the approximate formula given in Eq.(\\ref{gamma_app}) to compute the expression $t \\equiv t(r)$, instead of the exact solution of the Taub Eqs.(\\ref{Taub_Eq}), are clearly shown in Figs.~\\ref{eqts_comp_ad}--\\ref{eqts_comp_rad}. The EQTSs represented in these figures are computed at selected values of the detector arrival time both in the early ($\\sim 35$ s) and in the late ($\\sim 4$ day) phases of the afterglow. Both the fully radiative and fully adiabatic cases are examined. Note the approximate expression of the EQTS can only be defined for $\\gamma < \\gamma_d$ and $r > r_d$. Consequently, at $t_a^d=35$ s the approximate EQTSs are represented by arcs, markedly different from the exact solution (see the upper panels of Figs.~\\ref{eqts_comp_ad}--\\ref{eqts_comp_rad}). The same conclusion is found for the EQTS at $t_a^d=4$ days, where marked differences are found both for the fully radiative and adiabatic regimes (see the lower panels of Figs.~\\ref{eqts_comp_ad}--\\ref{eqts_comp_rad}). All the observational properties of GRBs, starting from the analysis of the prompt radiation \\citep{rbcfx02_letter}, to the luminosity in X- and $\\gamma$-ray bands, to their spectral distribution \\citep{Spectr1} as well as inferences on the possible presence or absence of beaming in GRBs, depend essentially on the structure of the EQTSs. In turn the determination of the EQTSs depends on the equations of motion of the baryonic pulse satisfying the Taub equations. The fact that the final results for the observable luminosity, spectral distribution, and substructures in the prompt radiation depend on $\\sim 10^8$ integration paths on different points on the EQTSs implies that the agreement between the theoretical predictions and the observations becomes a most stringent test for the validity of the equations of motion. The correct EQTSs are also essential for the identification of the energy source of the X and $\\gamma$ radiation in the GRB afterglow \\citep{rbcfx02_letter,Spectr1}. In conclusion, the approximate treatments largely overestimate (underestimate) the size of the EQTSs in the early (late) part of the afterglow. The theoretical slopes of the observables as a function of the arrival time \\citep[see, e.g.,][and references therein]{p99,p00,vpkw00} are therefore incorrectly evaluated. In the meantime, analytic expressions for the EQTSs have been obtained, validating the above results and allowing the theoretical estimate of the observables in GRB afterglows \\citep{br04}." }, "0403/astro-ph0403653_arXiv.txt": { "abstract": "{% In this paper we present a multifrequency and multiepoch study of \\object{PKS\\, 1502+106} at radio frequencies. The analysis is based on an EVN (European VLBI Network) dataset at 5 GHz and archive VLBA (Very Long Baseline Array) datasets at 2.3, 8.3, 24.4 and 43.1 GHz over a period of 8 years. The source is characterized by a multi--component one--sided jet at all epochs. The high--resolution images at 5, 8.3, 24.4 and 43.1 GHz show a curved jet morphology in the source. The radio core brightness temperature approaches the equipartition limit. Superluminal motions of $37.3\\pm9.3\\;c$, $22.0\\pm15.5\\;c$, $10.5\\pm2.6\\;c$ and $27.9\\pm7.0\\;c$ are measured in four distinct components of the jet. Our analysis supports the idea that the relativistic jet in \\object{PKS\\,1502+106} is characterised by extreme beaming and that its radio properties are similar to those of $\\gamma$--ray loud sources. ", "introduction": "\\begin{table*} \\centering \\caption{Logs of the observations} \\begin{tabular}{ccclc} \\hline Epoch &Freq &BW &Array and Available Telescopes$^{a}$&$D_{uv}^{b}$ \\\\ &(GHz) &(MHz) & &(km) \\\\ (1) & (2) & (3) & (4) &(5) \\\\\\hline 1994.52 &2.3 & 16 &VLBA(All 10) & 8600 \\\\ &8.3 & 16 &VLBA(All 10) & 8600 \\\\ 1997.85 &5.0 & 28 & EVN(Ef Sh Jb Ht Mc Nt On Tr Ur Wb) & $\\sim$10000 \\\\ 1998.11 &2.3 & 32 &VLBA(BR FD MK OV PT) GC KK & \\\\ & & &VLBA(BR FD HN KP LA NL OV PT SC) GC WF GN &5600 \\\\ &8.3 & 32 &VLBA(BR FD HN KP LA NL OV PT SC) GC WF GN &5600 \\\\ 2002.37 &24.4& 32 &VLBA(BR FD HN KP LA MK NL OV SC) & 8600 \\\\ &43.1& 32 &VLBA(BR FD HN LA MK NL OV SC) & 8600 \\\\ 2002.65 &24.4& 30 &VLBA(All 10) & 8600 \\\\ &43.1& 32 &VLBA(BR FD HN KP LA MK NL OV PT) & 6600 \\\\ 2002.99 &24.4& 64 &VLBA(FD HN KP LA MK NL OV PT SC) & 8600 \\\\ &43.1& 64 &VLBA(FD HN KP LA MK NL OV PT SC) & 8600 \\\\ \\hline\\end{tabular} \\label{obs} \\\\[0.3cm] \\raggedright $^{a}$ EVN telescope codes: Ef: Effelsberg, Sh: Shanghai, Ht: HartRAO, Jb: Jodrell, Mc: Medicina, Nt: Noto, On: Onsala, Tr: Torun, Ur: Urumqi, Wb: Westerbork Array; the VLBA observations at epoch 1998.11 are performed using subarrays made up of the VLBA antennas together with 4 geodetic antennas: GC (Fairbanks, AK USA), WF (Westford, MA USA), GN (Green Bank, WV USA) and KK (Kokee Park, HI USA);\\\\ $^{b}$ the longest baseline of the array, in unit of kilometer. \\end{table*} \\begin{table*} \\centering \\caption{Parameters of the images} \\begin{tabular}{lccllccc} \\hline Figure& Epoch & Freq & Real Beam$^{a}$ & Restored Beam$^{b}$ &S$_{Peak}~^{c}$ & rms$^{d}$ &Contours\\\\ & & (GHz) &Maj$\\times$Min(mas),P.A.($\\degr$)& Maj$\\times$Min(mas),P.A.($\\degr$)&(Jy/b) & (mJy/b)&(mJy/b)\\\\ (1) & (2) & (3) & (4) &(5) & (6) & (7) &(8) \\\\\\hline Fig.\\ref{fig1}a&1994.52 &2.3 & 7.18$\\times$3.78, $-1.67$&7.18$\\times$3.78, 0& 1.81&1.1 & 3.5$\\times$(-1,1,2,...,256)\\\\ Fig.\\ref{fig1}b&1998.11 &2.3 & 5.94$\\times$4.54, 13.7 &7.18$\\times$3.78, 0& 1.23&0.9 & 2.8$\\times$(-1,1,2,...,256)\\\\ Fig.\\ref{fig2}a&1997.85 &5.0 & 1.36$\\times$1.21, 64.7 &1.25$\\times$1.25, 0& 0.80&0.7 & 3.0$\\times$(-1,1,2,...,256)\\\\ Fig.\\ref{fig2}b&1994.52 &8.3 & 1.95$\\times$1.03, $-1.19$&1.25$\\times$1.25, 0& 1.37&0.9 & 2.8$\\times$(-1,1,2,...,256)\\\\ Fig.\\ref{fig2}c&1998.11 &8.3 & 1.62$\\times$1.23, 16.7 &1.25$\\times$1.25, 0& 0.68&0.6 & 2.0$\\times$(-1,1,2,...,256)\\\\ Fig.\\ref{fig3}a&2002.37 &24.4& 0.64$\\times$0.28, $-$1.08&0.64$\\times$0.28, 0& 1.26&0.8 & 2.1$\\times$(-1,1,2,...,256)\\\\ Fig.\\ref{fig3}b&2002.65 &24.4& 0.70$\\times$0.32, $-$3.77&0.64$\\times$0.28, 0& 0.73&0.9 & 2.8$\\times$(-1,1,2,...,256)\\\\ Fig.\\ref{fig3}c&2002.99 &24.4& 0.76$\\times$0.28, $-$6.0 &0.64$\\times$0.28, 0& 0.95&1.0 & 3.0$\\times$(-1,1,2,...,128)\\\\ Fig.\\ref{fig4}a&2002.37 &43.1& 0.37$\\times$0.16, $-$3.35&0.37$\\times$0.16, 0& 0.82&1.0 & 3.0$\\times$(-1,1,2,...,256)\\\\ Fig.\\ref{fig4}b&2002.65 &43.1& 0.66$\\times$0.24, $-$20.3&0.37$\\times$0.16, 0& 0.50&1.2 & 3.7$\\times$(-1,1,2,...,128)\\\\ Fig.\\ref{fig4}c&2002.99 &43.1& 0.44$\\times$0.16, $-$6.41&0.37$\\times$0.16, 0& 0.59&1.0 & 3.2$\\times$(-1,1,2,...,128)\\\\ \\hline \\end{tabular}\\label{image} \\\\[0.3cm] \\raggedright $^a$ the beam size directly measured from the visibilities (FWHM);\\\\ $^b$ the restored beam shown in the images (FWHM);\\\\ $^c$ the peak brightness in the image; \\\\ $^d$ the off--source rms noise level in the images. \\end{table*} One of the most significant observational results of extragalactic $\\gamma$--ray active galactic nuclei (AGNs) is that all EGRET--identified objects are radio--loud sources (Mattox et al. \\cite{Mattox}). Relativistic beaming in the jet is used to explain the EGRET identification in radio--loud AGNs. The EGRET--identified sources have on average much faster apparent superluminal motions than the general population of radio--loud sources (Jorstad et al. \\cite{Jorstad}). From a statistical analysis of $\\Delta PA$ (position angle differences between parsec-- and kiloparsec--scale structures) of EGRET--identified AGNs, Hong et al. (\\cite{Hong}) concluded that $\\gamma$--ray loud radio quasars typically show aligned morphologies on parsec and kiloparsec scales. It is still a matter of debate if the $\\gamma$--ray emission in AGNs is related to higher beaming in these sources. The radio--loud active galactic nucleus (AGN) \\object{PKS\\,1502+106} (4C 10.39, OR103), $z=1.833$ (Fomalont et al. \\cite{Fomalont}), is a highly polarized quasar (Tabara \\& Inoue \\cite{Tabara}). A high and variable degree of polarisation in the optical band is reported by Impey \\& Tapia (1988). It is known to be active and variable at radio, optical and X--ray wavelengths. In particular, the source exhibits intensity variations by a factor of 3 $\\sim$ 5 on timescales from weeks to months in the radio band; it shows intensity variations in the optical band with $m_{\\nu}$ ranging from 19.5 to 18.6, and in the X--ray band by a factor $\\geq 2$ at 1~keV (George et al. \\cite{George} and references therein). \\object{PKS\\, 1502+106} exhibits a 'core--jet--lobe' structure at radio wavelengths. A Very Large Array (VLA) image at 1.64 GHz (Murphy, Browne \\& Perley \\cite{Murphy}) shows that a continuous jet extends to the southeast and leads to a lobe located $\\sim$~7 arcsecond from the core. Very Long Baseline Interferometry (VLBI) observations (Fey, Clegg \\& Fomalont \\cite{Fey}; Fomalont et al. \\cite{Fomalont}; Zensus et al. \\cite{Zensus}) show a well--defined jet starting to the southeast and sharply bending to the east at a distance of 3--4 mas from the core. Our interest in \\object{PKS\\,1502+106} is related to the misalignment of the pc-- and kpc--scale radio structure in AGNs and its relation to the $\\gamma$--ray emission. Fichtel et al. (\\cite{Fichtel}) reported an EGRET flux density upper limit in PKS 1502+106 of $7 \\times 10^{-8}$ photons cm$^{-2}$ s$^{-1}$ from the Phase I results. However, the source was not detected in the following EGRET observations (Thompson et al. \\cite{Thompson}; Hartman et al. \\cite{Hartman}). To study the relation between the $\\Delta PA$ and $\\gamma$--ray emission, we observed a sample of $\\gamma$--ray blazars with the EVN at 5 GHz in Nov. 1997, in which PKS 1502+106 was observed as a gamma--ray source candidate (Hong et al. \\cite{Hong04}). In the next section we present the EVN observation of \\object{PKS\\,1502+106} at 5 GHz, and the data reduction. The analysis of the 5 GHz EVN image is performed in Sect. 3, where we also re--analyzed five epochs of Very Long Baseline Array (VLBA) datasets at 2.3, 8.3, 24.4 and 43.1 GHz obtained from the public archive of Radio Reference Frame Database (RRFID) for a consistent study with our data. To investigate the relativistic beaming in this source, we estimate the core physical parameters and measure the jet proper motions in Sect. 4. Conclusions and summary are given in Sect. 5. $H_0$= 65 km\\, s$^{-1}$Mpc$^{-1}$ and $q_0$=0.5 are used in this paper. \\begin{figure}\\centering \\resizebox{6.5cm}{!}{\\includegraphics{fig-1.eps}} \\vspace{-1mm} \\caption{VLBA images of \\object{PKS\\, 1502+106} at 2.3 GHz. A: epoch 1994.52; B: epoch 1998.11.} \\label{fig1} \\end{figure} \\begin{figure}\\centering \\resizebox{6.5cm}{!}{\\includegraphics{fig-2.eps}}\\vspace{-1mm} \\caption{VLBI images of \\object{PKS\\, 1502+106}. A: EVN image at 5 GHz at epoch 1997.85; B: VLBA image of at 8.3 GHz at epoch 1994.52; C: VLBA image at 8.3 GHz at epoch 1998.11. } \\label{fig2} \\end{figure} \\begin{figure}\\centering \\vspace{3mm} \\resizebox{6.5cm}{!}{\\includegraphics{fig-3.eps}} \\vspace{-1mm} \\caption{VLBA images of \\object{PKS\\, 1502+106} at 24.4 GHz. A: epoch 2002.37; B: epoch 2002.65; C: epoch 2002.99.} \\label{fig3} \\end{figure} \\begin{figure}\\centering \\vspace{3mm} \\resizebox{6.5cm}{!}{\\includegraphics{fig-4.eps}} \\vspace{-1mm} \\caption{VLBA images of \\object{PKS\\, 1502+106} at 43.1 GHz. A: epoch 2002.37; B: epoch 2002.65; C: epoch 2002.99.} \\label{fig4} \\end{figure} ", "conclusions": "In this paper we carried out a multifrequency and multiepoch study of \\object{PKS\\, 1502+106} at VLBI resolution. The source morphology is highly variable and the jet structure is very complex on this scale. We analyzed the structure of the jet, as seen in projection on the sky, and the changes in the position angle of the various jet components with the radial distance from the core. The results suggest that the jet components in \\object{PKS\\, 1502+106} trace the same curved path. The jet undergoes two major bends, the first implies a change in the P.A from $\\sim70\\degr$ to $\\sim110\\degr$ within 0.5 mas, while in the second bend the P.A. changes from $\\sim130\\degr$ to P.A.$\\sim80\\degr$ at 3--4 mas. Beyond that distance the VLBI jet points to the east at 7--15 mas. Based on the model fitting results on the VLBI core, we obtain a weighted mean value of the Doppler factor, $\\delta=4.5\\pm1.9$. The radio core brightness temperature in the source rest frame, T$_r = (2.0\\pm0.5) \\times10^{11}$K, approaches the equipartition limit. We detect superluminal motion in four components along the jet. The derived apparent speeds range between $10.5\\pm2.6\\,c$ and $37.3\\pm9.3\\,c$. Doppler boosting plays a major role in determining the observed properties of the source. In particular, the apparent speeds we derive suggest that the source is viewed under an angle to the line of sight $\\theta < 5^{\\circ}$, and that the bulk flow velocity is $\\beta_{intr}$ is $\\sim$ $0.999\\,c$. The $\\Delta PA$ between the pc-- and kpc--scale structure is about 30\\degr{}, indicating that PKS\\,1502+106 belongs to the aligned population. The superluminal speeds in 1502+106 are much higher than the average value in radio loud quasars. We therefore conclude that PKS\\,1502+106 is more beamed than the overall population of radio loud quasars, and that its radio properties are more similar to the $\\gamma$--ray loud quasars, although it is unclear if \\object{PKS\\, 1502+106} is a $\\gamma$--ray loud source to date. In particular, the superluminal speeds in this source are in the range found for the high redshift $\\gamma$--ray loud quasars in other high frequency radio surveys (Jorstad et al. 2001). A confirmation of $\\gamma$--ray emission from this source would be highly valuable for our understanding of the $\\gamma$--ray loudness phenomenon in radio loud quasars." }, "0403/hep-ph0403199_arXiv.txt": { "abstract": "Flat directions in the minimal supersymmetric standard model are known to deform into non-topological solitons, $Q$-balls, which generally possess both baryon and lepton asymmetries. We investigate how $Q$-balls evolve if some of the constituent fields of the flat direction decay into light species. It is found that the $Q$-balls takes a new configuration whose energy per charge slightly increases due to the decay. Specifically, we show that all the stable $Q$-balls eventually transform into pure $B$-balls via the decay into neutrinos. ", "introduction": "\\label{sec:intro} Non-topological solitons, such as $Q$-balls~\\cite{Coleman:1985ki} and $I$-balls~\\cite{Kasuya:2002zs}, play important roles in the particle cosmology, partly because they are (quasi-)stable objects. Their stability comes from the conservation law. The $Q$-ball is composed of a complex scalar field with an $U(1)$ charge, the conservation of which enables the $Q$-ball to be stable or long-lived. Similarly, $I$-balls are composed of a real scalar field (or a complex scalar field in a nearly straight-line motion), whose dynamics conserve the adiabatic charge. Since we can only see either these objects in their final state or their decay products in the present universe, it is important to study their evolution during the course of the expanding universe. The baryon asymmetry is one of the key observational results to uncover the history of the universe. Among many baryogenesis scenarios proposed so far, the mechanism proposed by Affleck and Dine~\\cite{Affleck:1984fy} can be realized in the minimal extension of the standard model, that is, the minimal supersymmetric standard model (MSSM). The mechanism makes use of one of flat directions, along which there is no classical potential in the supersymmetric limit. Although the flat direction is parametrized by a set of chiral superfields, it is convenient and even sufficient in the usual situation to express it in terms of a single complex scalar field dubbed `Affleck-Dine (AD) field,' $\\Phi$, which generally has nonzero baryon and lepton charges. The dynamics of the AD field not only generates the desired baryon asymmetry, but allows non-topological solitons to be formed. The properties of $Q$-balls in the context of the AD mechanism have been extensively studied~\\cite{Kusenko:1997si,Enqvist:1998en,Kasuya:1999wu, Kasuya:2000wx}. $Q$-balls can be stable or unstable depending on the situation. Stable $Q$-balls can be dark matter~\\cite{Kusenko:1997si,Kasuya:2000sc, Kasuya:2001hg}, which might solve the coincidence problem between the abundance of baryons and that of dark matter. Unstable but long-lived $Q$-balls can keep the baryon and lepton asymmetries stored inside them from being subject to the sphaleron effects. For instance, they can successfully generate both large lepton asymmetry and small baryon asymmetry~\\cite{Kawasaki:2002hq}. It is even possible to realize the late-time baryon appearance after the relevant BBN epoch with use of $Q$-balls~\\cite{Ichikawa:2004pb}. $I$-balls are also formed in the AD mechanism as an excited state of $Q$-balls. In the examples mentioned above, the single-field parametrization is simple and valid. However, if we examine the evolution of $Q$-balls closely, we have to pay much more attention to the fact that the flat direction is actually composed of several scalar fields. Indeed, the slepton condensate always decays into a pair of neutrinos, while the squark condensate can decay into hadrons only if its energy per one baryon charge exceeds the nucleon or meson masses. Therefore, if a $Q$-ball has nonzero baryon and lepton asymmetries and its energy per unit charge is small enough, the leptonic component decays, leaving only the baryon asymmetry inside $Q$-balls ({\\it i.e.}, pure $B$-balls). However, since the $Q$-ball solution is obtained under the assumption that the $U(1)$ charge, $Q$, is conserved, it is not trivial how the $Q$-ball configuration changes during the course of the partial decay. Do $Q$-balls come apart? Or do they change their shape to satisfy the modified $Q$-ball solution? It is the purpose of this study to answer these questions. In this paper, we investigate how $Q$-balls evolve if some of the constituent fields decay into lighter species. In the next section, we review the flat direction of the MSSM and the $Q$-ball solution. In section~\\ref{sec:model} we study the partial decay of $Q$-balls with a toy model of the flat direction. The results obtained there are confirmed with use of numerical calculation presented in section~\\ref{sec:nume}. Finally we give discussions and conclusions in section~\\ref{sec:discuss}. ", "conclusions": "\\label{sec:discuss} So far we have presented the analytical and numerical study of the partially decaying $Q$-balls with use of a toy model of flat directions. Let us consider now the implication of the results presented above. We have assumed that only a part of constituent fields decay, but our arguments apply to more generic case where there is a hierarchy between the decay rates, say, $\\Gamma_1 \\gg \\Gamma_2$. The asymmetry of the $Q$-ball continuously changes during the course of the decay. Most of the flat directions in the MSSM are composed of both squarks and sleptons, therefore they have both baryon and lepton number. If such a flat direction deforms into $Q$-balls and they are stable, our argument can be applied since the slepton condensate can always decay into a pair of neutrinos through the exchange of the gauginos. All through the decay processes, the $Q$-balls sequentially transform themselves toward pure $B$-balls with no lepton number. In the usual scenario considered so far, it is a flat direction composed only of squarks that forms pure $B$-balls. Therefore our analysis indicates that the pure $B$-balls are generated for most of flat directions, as a result of the decay into neutrinos. Specifically, all the $Q$-ball dark matter must be pure $B$-balls. Interestingly enough, the resultant $B$-ball might become unstable due to the increase of its energy per unit charge, leading to a complete decay of the $Q$-ball. In connection with the transformation of $Q$-balls into $B$-balls, the leptogenesis might become possible even for a flat direction with $B-L =0$. Another interesting possibility is that the $I$-ball can be naturally altered to Q-ball, if one of the constituent fields decay into something else. In particular, the $I$-ball composed of a flat direction that includes a leptonic field is generally transformed into a $Q$-ball. In fact, since the $I$-ball is a quasi-stable object, its conversion to a $Q$-ball was not known. The results of this study affords some new perspectives on the decay or stabilization processes of non-topological solitons made of flat directions. In summary, we have investigated the evolution of $Q$-balls, assuming that some of the constituent scalar fields decay into light particles. With a toy model of the flat direction, we have obtained the analytical solution of the $Q$-ball in the final state, and found that the spatial shape of the $Q$-ball does not change, but the orbit of the scalar field inside the $Q$-ball does change in order to compensate for the decay. In particular, the energy per unit charge generally increases due to the partial decay, which might induce further decay of the remnant scalar fields. Also we have performed numerical calculations to confirm these results. Again we stress that $Q$-balls remain $Q$-balls throughout the partial decay, due to both the charge conservation of the remnant scalar fields and the $D$-flat condition. \\subsection*" }, "0403/astro-ph0403135_arXiv.txt": { "abstract": "We present a method for computing the 6-dimensional coarse-grained phase-space density $f(\\Bx,\\Bv)$ in an N-body system, and derive its distribution function $v(f)$. The method is based on Delaunay tessellation, where $v(f)$ is obtained with an effective fixed smoothing window over a wide $f$ range. The errors are estimated, and $v(f)$ is found to be insensitive to the sampling resolution or the simulation technique. We find that in gravitationally relaxed haloes built by hierarchical clustering, $v(f)$ is well approximated by a robust power law, $v(f) \\propto f^{-2.5 \\pm 0.05}$, over more than 4 decades in $f$, from its virial level to the numerical resolution limit. This is tested to be valid in the $\\Lambda$CDM cosmology for haloes with masses $10^9-10^{15}\\msun$, indicating insensitivity to the slope of the initial fluctuation power spectrum. By mapping the phase-space density in position space, we find that the high-$f$ end of $v(f)$ is dominated by the ``cold'' subhaloes rather than the parent-halo central region and its global spherical profile. The value of $f$ in subhaloes near the virial radius is typically $>100$ times higher than its value at the halo centre, and it decreases gradually from outside in toward its value at the halo centre. This seems to reflect phase mixing due to mergers and tidal effects involving puffing up and heating. The phase-space density can thus provide a sensitive tool for studying the evolution of subhaloes during the hierarchical buildup of haloes. It remains to be understood why the evolved substructure adds up to the actual universal power law of $v(f) \\propto f^{-2.5}$. It seems that this behaviour results from the hierarchical clustering process and is not a general result of violent relaxation. ", "introduction": "\\label{sec:intro} The standard paradigm assumes that dark-matter haloes are the basic entities in which luminous galaxies form and live. The haloes dominate the gravitational potential over a wide range of radii and they have a crucial role in determining the galaxy properties. While many of the systematic features of halo structure and kinematics have been revealed by $N$-body simulations, the origin of these features is still an open theoretical puzzle, despite the fact that they are due to simple Newtonian gravity. The halo \\emph{density profile} $\\rho(r)$ is an example for such a puzzle. It is found in the simulations to have a robust non-power-law shape (which we refer to in general as ``NFW\"), with a log slope $-3$ near the virial radius, flattening gradually toward $-1$ at about 1\\% of the virial radius, and perhaps flattening further at smaller radii. \\citep[e.g.][and references therein]{ref:Nav97,ref:Mo98,ref:Kly01, ref:Pow03,ref:Hay04}. This density profile is insensitive (or at most weakly sensitive) to parameters of the cosmological model and the initial fluctuation power spectrum \\citep[e.g.,][]{ref:Col96,ref:Nav97,ref:Sub00, ref:Ric03, ref:Col03, ref:Nav04} indicating that its origin is due to a robust relaxation process rather than specific initial conditions. In particular, violent relaxation \\citep{ref:Lyn67} may be involved in shaping up the density profile. In addition, secondary infall may be important in the outer regions \\citep{ref:Gun72, ref:Dek81, ref:Fil84, ref:Hof85, ref:Whi92} while some argue that resonances may have a role in the inner regions (\\citealt{ref:Wei02}; but see \\citealt{ref:Val03}). Nevertheless, there is no clear understanding for why the haloes actually pick up the particular density profile they have. The properties of the \\emph{velocity dispersion} tensor is another theoretical puzzle. The haloes tend to be rather round, with a velocity dispersion profile that is slightly rising at small radii and slightly falling at large radii but is rather flat overall \\citep{ref:Hus99a, ref:Hus99b}. The profile of the anisotropy parameter $\\beta(r)$, which is a measure of radial versus tangential velocities, indicates near isotropy at small radii, which gradually develops into more radial orbits at large radii \\citep{ref:Col96, ref:Hus99a, ref:Hus99b, ref:Col00a}. For a spherical system in equilibrium, the $\\sigma(r)$ and $\\beta(r)$ are related to $\\rho(r)$ via the Jeans equation, but it is not at all clear why the haloes in the simulations choose the characteristic $\\sigma(r)$ or $\\beta(r)$ that they actually do. An interesting attempt to address the origin of the halo profile has been made by \\citet{ref:Tay01}, who measured a poor-man phase-space density profile by $f_{\\rm TN}(r) =\\rho(r)/\\sigma(r)^3$, and found that it displays an approximate power-law behaviour, $f_{\\rm TN} \\propto r^{-1.87}$, over more than two decades in $r$. Using the spheri-symmetric Jeans equation, they showed that this power law permits a whole family of density profiles, whose limiting case is a profile similar to NFW, which asymptotically approaches a slope $-0.75$ as $r \\rarrow 0$. The general power-law shape of $f_{\\rm TN}(r)$ is confirmed in the simulated haloes described below. This scale-free behaviour of $f_{\\rm TN}(r)$ is intriguing, and it motivates further studies of halo structure by means of phase-space density. Simulations of the $\\Lambda$CDM cosmology also reveal a roughly self-similar \\emph{hierarchical clustering} process, where smaller building blocks accrete and merge into bigger haloes. At every moment, every halo contains a substructure of subhaloes on top of a smooth halo component that has been tidally stripped from an earlier generation of substructure. Some of the important dynamical processes involved in this hierarchical halo buildup are understood qualitatively. These include, for example, the dynamical friction which governs the decay of the satellite orbits, the tidal stripping of subhaloes due to the host halo potential, and the mergers and flyby interactions of the subhaloes among themselves. However, a complete theoretical understanding of how these processes work in detail, and how they combine to produce the halo structure and kinematics, is lacking. Attempts have been made to explain an inner density cusp using toy models of dynamical stripping and tidal effects during the halo buildup by mergers \\citep{ref:Syr98,ref:Nus99,ref:ddh03,ref:Dek03}. However, a similar halo density profile seems to be produced also in some of the simulations where substructure has been artificially suppressed \\citep{ref:Mo99b, ref:Hus99a, ref:Avila01, ref:Bul02, ref:Alvarez02}, indicating that the process responsible for the origin of this density profile might be a somewhat more general feature of gravity and not unique to the merger scenario. The issue of halo substructure has become timely both because of its relevance to observations and its implications on other major issues in galaxy formation. Tidal tails and streams associated with dwarf satellite galaxies have been observed in the haloes of the Milky Way and M31 \\citep{ref:Iba94, ref:Iba01a, ref:Iba01b}, and they start to allow detailed modelling of the halo history through the satellite orbits. Gravitational-lensing observations provide preliminary indications for the presence of substructure in haloes at the high level predicted by the dissipationless $\\Lambda$CDM scenario \\cite[e.g.][]{ref:Dal02} In contrast, the observed number density of dwarf galaxies seems to be significantly lower, thus posing a ``missing dwarf problem'' \\cite{ref:Kly99b, ref:Mo99a}. Also, the ``angular-momentum problem'' of disk galaxies \\citep[e.g.][]{ref:Nav00,ref:Bul01a} is probably associated with the evolution of substructure in haloes \\citep{ref:Mal02, ref:mds02}. While the dwarf and angular-momentum problems necessarily involve baryonic processes, understanding the gravitational evolution of substructure is clearly a key for solving them. In order to better understand the origin the various aspects of halo structure and buildup mentioned above, we make here a first attempt at addressing directly and in some detail the \\emph{phase-space} structure of dark-matter haloes. The fundamental quantity in the dynamical evolution of gravitating systems is the full, six-dimensional, coarse-grained, phase-space density $f(\\Bx,\\Bv)$, which intimately relates to the underlying Vlasov equation and lies behind any relaxation process that may give rise to the virialized halo structure \\citep[][ chapter 4]{ref:Bin87}. Ideally, one would have liked to compute it free of assumptions concerning spherical symmetry, isotropy, or any kind of equilibrium. However, computing densities in a six-dimensional space is a non-trivial challenge which requires simulations of a very broad dynamical range. The state-of-the-art N-body simulations, with more than million particles per halo, allow an attempt of this sort for the first time. We describe below a successful algorithm for measuring $f(\\Bx,\\Bv)$, and study its relevant properties including the associated systematic and random uncertainties. We then apply this algorithm to simulated virialized haloes in the $\\Lambda$CDM cosmology. We report in this paper two surprising new results. First, we discover that the volume distribution function of the phase-space density, $v(f)$, displays a universal scale-invariant \\emph{power-law} shape, valid in all virialized haloes that form by hierarchical clustering. Second, we realise that this power law is not directly related to the overall density profile, but is rather driven by the halo \\emph{substructure}. This implies that the phase-space density provides a useful tool for studying the hierarchical buildup of dark-matter haloes and the evolution of substructure in them. In \\se{vf} we introduce $f(\\Bx,\\Bv)$ and $v(f)$. In \\se{computing} we describe the method of computing $f$ and $v(f)$ from an $N$-body halo, and summarise its properties and the associated errors, which are addressed in more detail in Appendix~\\ref{sec:errors}. In \\se{universal} we present the universal power-law shape of $v(f)$ based on several different simulations, and demonstrate its robustness to the mass scale and simulation technique. In \\se{substructures} we display maps of phase-space density and show that the high-$f$ contributions to $v(f)$ come from substructures within the parent halo. In \\se{conc} we summarise and discuss our results and future work. ", "conclusions": "\\label{sec:conc} Using Delaunay tessellation, we developed a method for measuring the 6-dimensional coarse-grained phase-space density $f(\\Bx,\\Bv)$ in $N$-body systems. We focused, in particular, on measuring the phase-space volume distribution function, $v(f)$. We applied this technique to several simulated haloes of $\\sim 10^6$ particles, formed by hierarchical clustering in the standard $\\Lambda$CDM scenario, and obtained two striking new results. First, $v(f)$ is well described by a power law, $v(f) \\propto f^{-2.5 \\pm 0.05}$, over 3 to 5 decades in $f$. The power-law regime starts at an $f$ value which corresponds to the characteristic size of the virialized halo. It ends at an $f$ value which is determined by the dynamical resolution limit of the specific simulation. Therefore, the true power-law range may extend to $f \\rightarrow \\infty$. This power law seems to be insensitive to the halo mass in the range $10^9-10^{15}\\msun$, indicating insensitivity to the exact slope of the fluctuation power spectrum, as long as the haloes are built by hierarchical merging of clumps bottom up. Second, this power-law originates from substructures within the halo rather then the overall trend with radius. The substructure completely dominates the high-$f$ parts of the $v(f)$ distribution. The infalling clumps seem to phase-mix --- by puffing up, heating and stripping --- as their orbits decay from the virial radius inwards toward the halo centre and they melt into the halo smooth background. Our first worry is that these results could be numerical artifacts, or severely contaminated by such. Based on our error analysis and tests with mock datasets, we argue that the $v(f)$ measured by the DTFE algorithm genuinely reflects the true phase-space properties of the given $N$-body system over a broad range of $f$. The question is whether the phase mixing suffered by the subclumps as they approach the halo centre might be an artifact of numerical effects such as two-body relaxation, leading to underestimated inner densities and/or overestimated internal velocities. A similar effect has been pointed out using a one-dimensional toy model \\citep{ref:Bin03}. The apparent agreement between simulations run with different codes and different resolutions is encouraging. In order to specifically address the effect of few-body relaxation, we intend to run twice a simulation of the same halo with the same number of particles but with a different force resolution (ongoing work with F. Stoehr). Assuming that the simulations genuinely reflect the true physical behaviour under the Vlasov equation, the origin of the robust power-law shape of $v(f)$ from the merging substructure becomes a very interesting theoretical issue. As demonstrated in \\se{substructures}, a simple model using the mass function and the scaled profiles of the general halo population in the $\\Lambda$CDM scenario does not reproduce the correct power law. This, and the apparent trend of the $f$ spikes with radius, indicate that the structural and kinematic evolution of the subhaloes in the parent halo are important. Studies of tidal heating and stripping may be found useful in this modelling. It would be interesting to follow the phase-space evolution and the contribution to the overall $v(f)$ by a single, highly resolved subhalo, or many of those, as they orbit within the parent halo and approach its centre. This may help us understand the nature of the interaction between the parent halo and its subhaloes, and the origin of the $v(f)$ power law (ongoing works with E. Hayashi and with B. Moore). Another more general but speculative possibility is that the $f^{-2.5}$ power law represents some sort of a cascade of relaxation processed in phase-space, in which high phase-space densities turn into lower (coarse-grained) densities through the process of mixing. In general, the fact that our findings are expressed in terms of the fundamental concept of phase-space density should make them more directly accessible to analytical treatment. In this respect, it may prove beneficial to investigate more closely the time evolution of the $v(f)$ of a cosmological halo and its components. This may shed light on the connection between the $v(f)$ power-law behaviour and the relaxation processes within the halo. We saw that the power-law behavior of $v(f)$ is limited to the virial regime. It would be interesting to learn how this shape evolves in time as the halo virializes. A preliminary study (to be concluded and reported in another paper) indicates that in the intermediate-$f$ regime the $v(f)$ of a pre-virialized system is significantly flatter than $f^{-2.5}$, while in the high-$f$ regime it drops in a much steeper way. The $f^{-2.5}$ behaviour seems to be a feature unique to virialized systems. We learnt that in the haloes that are built by hierarchical clustering, the power-law behaviour $v(f)\\propto f^{-2.5}$ reflects the halo substructure. It would be interesting to find out whether this power-law behaviour actually requires substructure, or it is a more general phenomenon of virialized gravitating systems, valid independently of substructure. One way to answer this question would be to analyse simulated haloes in which all fluctuations of wavelengths smaller than the halo scale were removed, resulting in a smooth halo formed by monolithic collapse, with no apparent substructure in the final configuration. As described in \\se{intro}, such haloes are known to still have NFW-like density profiles in real space, and one wonders whether they also have the magic power-law $v(f)$. There are preliminary indications for a steeper $v(f)$ in this case (Arad, Dekel \\& Moore, in preparation). If confirmed, it would indicate that the $f^{-2.5}$ behaviour, while insensitive to the exact slope of the initial power spectrum, is unique to the hierarchical clustering process, and is not a general result of violent relaxation. Our current results are just first hints from what seems to be a promising rich new tool for analysing the dynamics and structure of virialized gravitating systems. The analysis could become even more interesting when applied to haloes including the associated gaseous and stellar components." }, "0403/astro-ph0403303_arXiv.txt": { "abstract": "{We present an X-ray study of the high metallicity young open cluster Blanco\\,1 based on {\\em XMM--Newton} data. X-ray spectroscopy of cluster members is presented for the first time as well as new X-ray distribution functions of late-type stars. We detected all known dF and dG stars in the EPIC field and 80\\% and 90\\% of dK and dM stars, respectively. The X-ray spectral analysis of the X-ray brightest cluster stars and X-ray color analysis of a larger sample show that a model with two temperatures (at about 0.3 and 1 keV) explains the quiescent activity phase spectra. We discuss also the nature of unidentified X-ray sources in the observed region and their X-ray spectral properties. ", "introduction": "% Open clusters are powerful laboratories to test the models of star formation and evolution as well as the metal enrichment in the Galaxy. In fact, they are naturally selected samples of stars with same age, composition and environmental formation conditions. In the last two decades, X-ray observations of open clusters have assumed great importance. Since the early '80s, the open cluster X-ray observations of first {\\em Einstein} and subsequently ROSAT, showed that young cluster stars are stronger X-ray sources than the Sun (\\citealp[see][]{Micela2002} and reference therein cited). The evolution of X-ray activity is correlated with the evolutionary angular momentum losses, older stars being less luminous than the younger ones, but other, more subtle, factors like pre-main-sequence history and chemical composition could also play a role. The comparison of the properties of different clusters is essential for an understanding of the importance of these factors. It is expected that stellar chemical composition influences the processes of coronal emission and activity: X-ray spectra of stars with high metal content are expected to have enhanced line emission contribution with respect to low metallicity stars. Moreover, the extent of the convective zone in solar type stars is also determined by metal content, thus resulting in a possible enhanced dynamo efficiency. Investigations of open clusters with different metallicity are of great importance for the study of the relation between stellar structure and X-ray activity. Blanco\\,1 is a young open cluster noticeably more distant from the Galactic Plane (about 240 pc) than the young open cluster scale height ($\\sim 100$ pc). Its age, around 100 Myr, is very similar to that of the Pleiades and NGC\\,2516 clusters \\citep{deEpst85,West88}, while its metallicity ([Fe/H]=+0.23 dex) is significantly higher than that of the Sun and the Pleiades \\citep{Edv95,Pan97,Jef99}. The high metallicity of Blanco\\,1 offers a suitable benchmark for the connection between stellar structure, activity and convection. Measurements of Li abundance by \\citet{Jef99} do not fit standard mixing models for stars of this age and metallicity, thus implying some Li-depletion inhibition. Our previous work (\\citealp{Giusi99}; \\citealp{bl1hri}, Paper I) based on ROSAT data, showed that the cluster dG and dK stars have an overall X-ray emission similar to that of the Pleiades while the dM stars in Blanco\\,1 appeared more luminous in the X-ray band with respect to the Pleiades and $\\alpha$\\,Per clusters. Prior to Paper I, membership catalogues for stars down to late K types were based only on photometry \\citep{deEpst85,Pan97}, while a few radial velocity measurements are given in \\citet{Jef99} and \\citet{Edv95}. In Paper I the membership of the cluster stars in the ROSAT field of view was defined by means of proper motion analysis and it was possible to derive X-ray luminosity distributions (XLDs) for dF, dG, dK and dM types. In that work, due to the limited sensitivity of ROSAT, the XLD of dM stars was significantly affected by the large number of upper limits. Furthermore, the lack of spectral capability of the ROSAT-HRI camera, did not allow us to derive any spectral features of the cluster coronae. The {\\em XMM-Newton} observation addresses these two issues: the large effective area of its three X-ray telescopes allows us to detect essentially all cluster members in the observed region and the moderate spectral resolution of the EPIC camera allows us to explore the main characteristics of the coronal spectra. In this work we present the X-ray spectroscopy and photometric analysis of the cluster region as follows: in Sect. 2 we describe the observations, the basic processing of the data and the source detection results. In Sect. 3 we discuss the X-ray spectral analysis of the cluster stars and their XLDs. In Sect. 4 we discuss the nature of the unidentified sources and in Sect. 5 we summarize our conclusions. ", "conclusions": "% We have presented our analysis of an {\\em XMM-Newton} observation of the young open cluster Blanco\\,1 addressing the issues of the completeness of detection among X-ray cluster sources and, for the first time, of low resolution X-ray spectroscopy of its members. We have detected 190 X-ray sources by using a wavelet based algorithm developed specifically for the EPIC camera. Detections have been derived on the combined MOS 1, 2 and {\\em p--n} datasets, making full use of the EPIC camera capability and improving the detection sensitivity. Of the 190 detected sources, 36 are identified with cluster members, while 5 cluster members remain undetected. As compared with previous ROSAT observations, the {\\it XMM-Newton} rate of detection has increased from 61\\% to 88\\%. Six cluster sources with more than 1000 counts have been the subject of spectral analysis. A 2-T model fits the main spectrum features well and allows us to identify a ``cool'' component around 0.3 keV and a ``hot'' component around 1 keV, both with similar emission measures. These values are in good agreement with the XMM description of coronal spectra for young active stars like the Pleiades. A 1-T model at $\\sim 0.6$ keV explains the soft spectrum of an intermediate type star (spectral type $\\sim$ A8) of the cluster. We consider that the X-ray emission is intrinsic to the A star itself and not due to an unresolved late companion. By means of two X-ray color indices, the X-ray spectral properties of a larger cluster sample were explored and found to be consistent with the above 2-T model. The X-ray luminosity distribution function of dM-type stars of the cluster indicates a probable difference in emission levels with respect to the Pleiades. Saturation of $\\log \\mathrm{L}_\\mathrm{X}/\\mathrm{L}_\\mathrm{bol}$ is not clearly evident down to mid-K stars, slightly later than in Pleiades. Appendix A lists non-cluster optical counterparts for 90 of the 190 X-ray sources, identified either with the GSC-II and USNO-B1 optical catalogs or with the 2MASS infrared catalog. One star is suggested as new probable low mass cluster member. Listed in Appendix B are 64 sources that remain unidentified: most of them show X-ray colors quite different from those of the cluster, suggesting a different nature. However, for two of them the color analysis in Sect. 4 suggests agreement of their spectrum with those of cluster coronae." }, "0403/astro-ph0403629_arXiv.txt": { "abstract": "We present results of an analysis of the structural and kinematical properties of a sample of elliptical-like objects (ELOs) identified in four hydrodynamical, self-consistent simulations run with the DEVA code (Serna et al. 2003). Star formation has been implemented in the code through a simple phenomenological parameterization, that takes into account stellar physics processes only implicitly through the values of a threshold gas density, $\\rho_{\\rm g,thres}$, and an efficiency parameter, $c_*$. The four simulations operate in the context of a $\\Lambda$CDM cosmological model consistent with observations, resolve ELO mass assembly at scales up to $\\simeq$ 2 kpc, and differ in the values of their star formation parameters. Stellar masses, projected half-mass radii and central l.o.s. velocity dispersions, $\\sigma_{\\rm los, 0}$, have been measured on the ELO sample and their values compared with data from the Sloan digital sky survey. For the first time in self-consistent simulations, a good degree of agreement has been shown, including the Faber-Jackson and the $D_n - \\sigma_{\\rm los, 0}$ relations, among others, but only when particular values of the $\\rho_{\\rm g,thres}$ and $c_*$ parameters are used. This demostrates the effect that the star formation parameterization has on the ELO mass distribution. Additionally, our results suggest that it is not strictly necessary, at the scales resolved in this work, to appeal to energy sources other than gravitational (as for example supernovae feedback effects) to account for the structure and kinematics of large ellipticals. ", "introduction": "One of the most challenging open problems in modern cosmology is the origin of the local galaxies of different Hubble types we observe to-day. Among them, ellipticals are the easiest to study. They form the most homogeneous family and show the most precise regularities in the form of correlations among some pairs of their observable parameters. The Sloan digital sky survey (SDSS, see York et al. 2000) sample of early-type galaxies, containing to date 9000 galaxies from different environments, provides a new standard of reference for nearby elliptical galaxies. Its analysis confirmed correlations previously established, such as those involving structural and kinematical parameters (the $L - \\sigma_{\\rm los, 0}$ or Faber-Jackson relation, 1976; the surface-brightness - effective radius relation, Kormendy 1977; the $D_n - \\sigma_{\\rm los, 0}$ relation, Dressler et al. 1987; among others, see Bernardi et al. 2003a, 2003b, 2003c, 2003d, and references quoted therein). These correlations, as well as the [$\\alpha/$Fe] ratio trend with $\\sigma_{\\rm los, 0}$ (Jorgensen 1999), demand short formation time-scales and old formation ages for the bulk of the stellar populations of ellipticals. These requirements are naturally met by one of the scenarios proposed so-far to explain galaxy formation and evolution: the so-called {\\it monolithic collapse} scenario (ellipticals would form at high $z$ in a single burst of star formation, and would passively evolve since then; Patridge \\& Peebles 1967; Larson 1974). The competing {\\it merging scenario} (galaxy mass assembly takes place gradually through repeated mergers of smaller subunits; Toomre 1977; Kauffmann 1996) meets some difficulties at explaining the correlations above as well as other observations on ellipticals, see Peebles 2002 and Matteucci 2003 for details and discussions. But the monolithic collapse scenario does not recover all the currently available observations on ellipticals either. Such are, for example, the range in ages their stellar populations span in some cases or their kinematical and dynamical peculiarities (Trager et al. 2000; Menanteau, Abraham \\& Ellis 2001), indicating that an important fraction of present-day ellipticals have recently experienced merger events. In order to reconcile all this observational background within a formation scenario, it is preferable to study galaxy assembly from simple physical principles and in connection with the global cosmological model. Self-consistent gravo-hydrodynamical simulations are a very convenient tool to work out this problem (Navarro \\& White 1994, Tissera, Lambas \\& Abadi 1997, Thacker \\& Couchman 2000). The method works as follows: initial conditions are set at high $z$ as a Montecarlo realization of the field of primordial fluctuations to a given cosmological model in a periodic, homogeneously sampled box; then the evolution of these fluctuations is numerically followed up to $z =0$ by means of a computing code that solves the N-body plus hydrodynamical evolution equations. In this way, the uncertainties resulting from prescriptions on dynamics and gas cooling and heating, present in other methods such as semi-analytical ones (Kauffmann et al. 1999, Mathis et al. 2002) can be removed, only star formation needs further modelling. Individual galaxy-like objects (GLOs) naturally appear as an output of the simulations, no prescriptions are needed as far as their mass assembly processes are concerned. Moreover, self-consistent simulations directly provide detailed information on each individual GLO at each $ z$, namely its six dimensional phase space structure, as well as the temperature and age distributions of its gaseous and stellar components, respectively. From this information, the parameters characterizing each GLO can be estimated and compared with observations (see, for example, S\\'aiz et al. 2001, concerning disk galaxies). The first step in the program of studying the origin of galaxies through self-consistent simulations, is to make sure that they form GLO samples of different Hubble types that have counterparts in the real local Universe. In particular, the possible simple correlations involving structural and kinematical parameters must be recovered. A detailed analysis of this kind was not yet available for ellipticals (see previous work in Kobayashi 2002; Sommer-Larsen, Gotz, \\& Portinari 2002; Meza et al. 2003). We present in this paper the results of an analysis of the structure and kinematics of a sample of elliptical-like objects (ELOs) identified in four self-consistent hydrodynamical simulations run in the framework of a flat $\\Lambda$CDM model consistent with observations. We have used DEVA, an AP3M-SPH code particularly designed to study galaxy assembly in a cosmological context. In DEVA, special attention has been paid that the implementation of conservation laws (energy, entropy and angular momentum) is as accurate as possible (see Serna, Dom{\\'{\\i}}nguez-Tenreiro, \\& S\\'aiz, 2003 for details, in particular for a discussion on the observational implications of violating some conservation laws). Star formation (SF) processes have been implemented in the code through a simple parameterization, similar to that first used by Katz (1992), that includes a threshold gas density, $\\rho_{\\rm g,thres}$ and an efficiency parameter, $c_{*}$, determining the SF timescales according with a Kennicutt-Schmidt law\\footnote{See Elmegreen (2003) for a discussion on the possibility that this law can be explained as a result of SF processes at the scale of molecular cloud cores, through an interstellar medium (ISM) gas structure whose density, prior to SF, can be described by a log-normal probability distribution, as Wada \\& Norman (2001) found in their simulations } (Kennicutt 1998). Galaxy-like objects of different morphologies have been identified in the simulations (S\\'aiz, Dom\\'{\\i}nguez-Tenreiro \\& Serna 2002; S\\'aiz 2003). The aim of this paper is to show that some of those identified as ELOs, have, at a structural and kinematical level, counterparts in the local Universe, including parameter correlations. Data have been taken from the SDSS as analyzed by Bernardi et al.\\ (2003a, 2003b), Kauffmann et al. (2003a, 2003b) and Shen et al., 2003. A brief account on ELO assembly is as follows: it mainly occurs through a multiclump collapse at rather high $z$ involving many clumps; collapse takes the clumps closer and closer along filaments causing them to merge at very low relative angular momentum and, consequently, at short timescales. This results into fast SF bursts at high $z$ that transform most of the available gas into stars. The frequency of head-on mergers decreases with $z$. ELO stellar populations are mostly old, and a trend exists with $\\sigma_{\\rm los, 0}$, as suggested by some observations (Thomas, Maraston \\& Bender, 2002). ", "conclusions": "The degree of consistency reported here between sizes, velocity dispersions and stellar masses of simulated ELOs and SDSS data is very good. The agreement is particularly outstanding when the simplicity of our working scheme is recalled: ELO assembly has been simulated in the context of a cosmological model roughly consistent with observations; Newton laws and hydrodynamical equations have been integrated in this context, with a standard cooling algorithm and a SF parameterization through a Kennicutt-Schmidt-like law, containing our ignorance about its details at sub-kpc scales. No further hypotheses to model the assembly processes have been made. Our results suggest that, at least for the more massive objects, say $\\ge 6 \\times 10^{10}$ M$_{\\odot}$, it is not strictly necessary to appeal to energy sources, {\\it at the scales resolved in this work}, that is, up to $\\sim 2$ kpc, other than gravitational to account for their sizes, velocity dispersions, stellar masses and their correlations. This result is consistent with the idea that SNe explosions are not relevant at these scales, but only at smaller scales, to make shells and trigger star formation in molecular cloud cores. Their effect would be accounted for in the SF parameters (Elmegreen 2003). This work strongly suggests that the SF parameterization is a key ingredient to determine the compactness of elliptical galaxies. A good agreement has been found in the correlations addressed here with SDSS data, but in the case of those involving sizes, only when particular values of the SF parameters are used. Our results push the problem of elliptical galaxy formation from understanding their mass assembly at scales $>$ kpc in a cosmological context (Dom\\'{\\i}nguez-Tenreiro et al. 2003, in preparation), to understanding how SF was regulated at high $z$ as the gas falls within collapsing volumes, or other shock locations, so as to proceed just with the efficiency necessary to produce the sizes observed in to-day ellipticals. It is a pleasure to thank J. Silk and J. Sommer-Larsen for useful information on the topics addressed in this paper. This work was partially supported by the MCyT (Spain) through grant AYA-0973 from the Programa Nacional de Astronom\\'{\\i}a y Astrof\\'{\\i}sica. We also thank the Centro de Computaci\\'on Cient\\'{\\i}fica (UAM, Spain) for computing facilities. \\clearpage" }, "0403/astro-ph0403073_arXiv.txt": { "abstract": "The large scale anisotropies of WMAP data have attracted a lot of attention and have been a source of controversy, with many of favourite cosmological models being apparently disfavoured by the power spectrum estimates at low $\\ell$. All of the existing analyses of theoretical models are based on approximations for the likelihood function, which are likely to be inaccurate on large scales. Here we present exact evaluations of the likelihood of the low multipoles by direct inversion of the theoretical covariance matrix for low resolution WMAP maps. We project out the unwanted galactic contaminants using the WMAP derived maps of these foregrounds. This improves over the template based foreground subtraction used in the original analysis, which can remove some of the cosmological signal and may lead to a suppression of power. As a result we find an increase in power at low multipoles. For the quadrupole the maximum likelihood values are rather uncertain and vary between 140-220$\\mu {\\rm K}^2$. On the other hand, the probability distribution away from the peak is robust and, assuming a uniform prior between 0 and $2000\\mu {\\rm K}^2$, the probability of having the true value above $1200\\mu {\\rm K}^2$ (as predicted by the simplest $\\Lambda CDM$ model) is 10\\%, a factor of 2.5 higher than predicted by WMAP likelihood code. We do not find the correlation function to be unusual beyond the low quadrupole value. We develop a fast likelihood evaluation routine that can be used instead of WMAP routines for low $\\ell$ values. We apply it to the Markov Chain Monte Carlo analysis to compare the cosmological parameters between the two cases. The new analysis of WMAP either alone or jointly with SDSS and VSA reduces the evidence for running to less than 1-$\\sigma$, giving $\\alpha_s=-0.022\\pm 0.033$ for the combined case. The new analysis prefers about 1-$\\sigma$ lower value of $\\Omega_m$, a consequence of an increased ISW contribution required by the increase in the spectrum at low $\\ell$. These results suggest that the details of foreground removal and full likelihood analysis are important for the parameter estimation from WMAP data. They are robust in the sense that they do not change significantly with frequency, mask or details of foreground template marginalization. The marginalization approach presented here is the most conservative method to remove the foregrounds and should be particularly useful in the analysis of polarization, where foreground contamination may be much more severe. ", "introduction": "Data analysis of cosmic microwave background maps is a challenging numerical problem. The question that we want to answer is the probability (or likelihood) of a theoretical model given the data. In order to evaluate the exact likelihood of a theoretical power spectrum of CMB fluctuations given a sky map of these fluctuations it is necessary to invert the theoretical covariance matrix. This operation scales as $O(N^3)$, where $N$ is the length of the data vector and is currently limited by practically available computer technology to $N \\lesssim 10^4$. One is hence forced to use approximate estimators when inferring the power spectrum from data such as WMAP satellite \\citep{2003ApJS..148....1B}, which have 1-2 orders of magnitude more independent measurements. The most popular methods are the pseudo-Cl (PCL) method (see e.g. \\citep{2002ApJ...567....2H}) and the Quadratic Maximum Likelihood (QML) estimator (see e.g. \\citep{1997PhRvD..55.5895T}). Both of these methods produce as an intermediate step estimates of multipole moments $C_\\ell$ and approximate methods have been developed to describe their probability distributions as accurately as possible \\citep{2000ApJ...533...19B,2003ApJS..148..195V}. These perform satisfactorily for high $\\ell$ values, where the central limit theorem guarantees a Gaussian distribution (in offset lognormal transformed variables) will be a good approximation. Unfortunately, these methods are much less reliable at low multipoles, where the distributions are not Gaussian. The situation is complicated further by the masks applied to the data to remove the galactic foreground contamination and by the marginalization of unwanted components, all of which makes analytic approach unreliable. In \\cite{2003astro.ph..7515E} it was suggested to use a hybrid approach using QML on degraded maps at low $\\ell$ and PCL at higher multipoles. The issue of the exact values of multipole moments in WMAP data has attracted much attention since the original analysis by WMAP team \\cite{2003ApJS..148..175S}. Several unusual features have been pointed out already in the original analysis. One of these was the correlation function, which appears to almost vanish above $60^{\\circ}$. Another was the low value of the quadrupole. With the PCL analysis the value of the quadrupole was found to be $\\sim 123 \\mu {\\rm K}^2$, compared to the expected value of $\\sim 1200 \\mu {\\rm K}^2$ for the simplest $\\Lambda CDM$ model. The probability for this low value was estimated to be below 1\\%, depending on the parameter space of models. The discussion of the statistical significance of the low values of quadrupole and octopole in the WMAP data \\citep{2003MNRAS.346L..26E,2003JCAP...07..002C,2003JCAP...09..010C,PhLB..570..145L} has sparked a renewed interest into the so called estimator induced variance \\citep{2004MNRAS.348..885E} - the error in the likelihood evaluation arising due to the use of an estimator rather than the exact expression. In \\cite{2004MNRAS.348..885E} it has been argued that [QML estimator performs significantly better than the PCL estimator and that the true value of the quadrupole probably lies in the range around $170-250 \\mu {\\rm K}^2$. However, only the maximum likelihood value was computed and not the full likelihood distributions so the statistical significance of this result and its effect on the parameter estimation remained unclear. In addition, the role of foregrounds and monopole/dipole removal has not been explored in detail. In this paper we take a different approach. We argue that the actual value of the best fitted quadrupole (and other multipoles) is not of the main interest, since it can be quite sensitive to the details of the foreground subtraction procedure, type of mask used and numerical details of the analysis (in fact, the various values proposed so far may even be statistically indistinguishable if the likelihood function at the peak is very broad). What is more important is the probability or likelihood of a model given the data, compared to another model that may, for example, fit the data better. This is encapsulated in the likelihood ratio between models and within the Bayesian context is the only information we really need to asses the viability of cosmological models that belong to a certain class. In this paper we perform the exact likelihood calculation by a direct inversion of the covariance matrix for the low resolution maps, thus eliminating all the uncertainties related to estimator variance approximations. Since we use low resolution maps with less than 3000 pixels we can do the inversions with a brute force linear algebra routines. This means we cannot do the analysis on all of the multipole moments, so we analyse low multipoles with the exact method and use PCL analysis for the higher multipoles, where the two methods agree with each other and where the approximate variance estimates developed for PCL analysis are likely to be valid. Second issue we wish to address in this paper is the question of foreground subtraction. This is done in two steps. First, pixels with high degree of contamination are completely removed from the data. This results in the so called KP2 (less aggressive, 85\\% of the sky) and KP0 (more aggressive, 75\\% of the sky) masks \\citep{2003ApJS..148...97B}. There remains contamination even outside these masks in individual frequency channels. This contamination can be further reduced using templates and/or frequency information \\citep{2003ApJS..148...97B}. In WMAP analysis the templates were fitted for and subtracted out of WMAP data. Even with a perfect template there is a danger that this procedure can oversubtract the foregrounds, since one is essentially subtracting out the maximum amplitude consistent with the template which could include some of the signal. Instead, here we do not subtract out the templates, but marginalise over them by not using any information in the data that correlates with a given template. This procedure has not been applied to WMAP data in previous analyses. It guarantees that there is no statistical bias caused by the foreground removal. Some of the templates that were subtracted in WMAP analysis, particularly 408MHz Haslam synchrotron radiation map \\citep{1982A&AS...47....1H}, are of poor quality. WMAP produced a better set of templates applying Maximum Entropy Method (MEM) to WMAP maps in several frequency channels using templates as priors only \\citep{2003ApJS..148...97B}. In addition to the Haslam synchrotron map, they used \\cite{2003ApJS..146..407F} H-$\\alpha$ map as a tracer of free-free emission and the SFD dust template based on \\cite{1998ApJ...500..525S}. This process resulted in three MEM derived foreground maps. These, however, were not used to infer the power spectrum. Instead, the official power spectrum was determined from the integrated single frequency maps and the same templates that were used as priors for the MEM map making procedure, ignoring the MEM derived maps. The MEM derived maps are likely to be the most faithful representation of the foregrounds. When used in power spectrum inference, however, they must be used with care due to complicated nature of their signal and noise correlations \\citep{2003ApJS..148...97B}. Nevertheless, on the largest scales, where receiver noise is negligible, they are probably the best available option. We therefore use the integrated single channel maps and the MEM derived foreground templates as a basis of our work. Note that in foreground marginalization procedure no template is actually removed from the data and there is no danger of introducing noise correlations that could significantly affect the power spectrum estimates. We perform this process on foreground unsubtracted maps of the V and W channels of the WMAP satellite. We use both KP2 and KP0 masks and project out the remaining galactic contamination using MEM inferred maps of dust, synchrotron and free-free foregrounds. We use the likelihood evaluated in this way to asses the statistical significance of departure from the concordant model at low multipoles and to perform the statistical analysis of cosmological models given the data. WMAP team also produced the so called \\emph{Internal Linear Combination} (ILC) map of the CMB emission, by using internal maps at various frequencies to decompose them into CMB and foreground components. This approach is not based on any templates and so uses less information than in principle available. While visually these maps appear to be relatively free of contamination outside the galactic plane, there are still artifacts within the plane. This means that one must be careful when projecting out monopole and dipole: one should not simply remove them from the all-sky map, since they could be contaminated by galactic emission at the canter and this would leave a residual offset outside the galactic plane, which could contaminate all of low multipoles. One must again apply the marginalisation over monopole and quadrupole on the masked map to eliminate any contamination in the final result. A similar approach has been taken by \\cite{2003PhRvD..68l3523T} and \\cite{2004astro.ph..3098E}, who produced their own versions of ILC maps. Since we argue that the best method is to use single frequency maps together with correct templates and we use ILC map for illustration and cross-check purposes only, we do not consider these alternative ILC solutions further. ", "conclusions": "In this paper we have developed routines to calculate the exact likelihood of the low resolution WMAP data. We have projected out unwanted foreground components by adding the foreground templates to our covariance matrix with large variance. Both of these methods have not been applied to WMAP data before and should improve upon the existing analyses. We have tested the robustness of our results by applying the method to many different combinations of observing frequency, mask, smoothing and templates and found consistent results among these various cases. In particular, we find consistent results if we marginalize only over dust in $W$ channel as opposed to all 3 foreground templates, if we use templates external to WMAP instead of WMAP MEM templates, if we use KP0 instead of KP2 mask, if we use ILC maps instead of individual V or W frequencies or if we use Healpix windows instead of gaussian smoothing. The two most important features of our procedure are thus marginalization over dust and exact likelihood analysis. Important differences exist between our results and previous work. We find higher values of the lowest multipoles, which is partly a consequence of template subtraction method used in WMAP analysis. This procedure would certainly remove some of the real power, although it is difficult to estimate how much and the differences could also be just a statistical fluctuation. For the maximum likelihood value of the quadrupole we find values between the original WMAP analysis and subsequent reanalysis by \\cite{2004MNRAS.348..885E}. The differences are within the estimated error of the foreground contamination and we argue that the actual value is not very reliable given how broad the likelihood is at the peak. More important is the shape of the likelihood function, which we find to be broader than in the WMAP team provided likelihood evaluation, which underestimates the errors compared to our analysis. This lowers the statistical significance of the departure of the data from the concordant model. Within a Bayesian context and assuming a flat prior on the distribution of quadrupoles we find the probability that a model exceeds the concordance model predicted quadrupole to be 10\\%. We also do not find anything particularly unusual in the correlation function and in the joint quadrupole-octopole analysis. We combine the full likelihood calculation with foreground marginalization at low $\\ell$ with the original WMAP PCL analysis at high $\\ell$ to generate Monte Carlo Markov Chains, whose distribution converges to the probability distribution of theoretical models given the data and assumed priors. The main effect of the new analysis is on the running of the spectral index, for which the marginal 2 sigma evidence for $\\alpha_s<0$ present in the original analysis and in the recent analysis of WMAP+VSA \\cite{2004astro.ph..2466R} (see also \\cite{2004astro.ph..2359R}) is reduced to below 1 sigma. Using the exact WMAP likelihood analysis will be essential for attempts to determine the running of the spectral index by combining WMAP with either the small scale CMB data or with the upcoming Ly-$\\alpha$ forest analysis from SDSS. In all of these cases the exact method increases the value of the running by pushing up the CMB spectrum at large scales. Another parameter which is significantly affected is the matter density $\\Omega_m$ or, equivalently, the dark energy density $\\Omega_{\\Lambda}$. We find $\\Omega_m$ to be reduced by the new analysis because of the added power at low multipoles, which is most easily accounted for by an increase in ISW contribution. We have shown that the effects of the improved likelihood analysis presented here can be significant for the determination of cosmological parameters. We expect the methods applied here will be equally important for the analysis of polarization data in WMAP, where the foregrounds play a much more important role and where a full likelihood analysis of joint temperature and polarization data is necessary to extract the maximum amount of information. Current analysis of temperature-polarization data is rather unsatisfactory, since it is based on the cross-spectrum information alone. Without having access to the full polarization maps we cannot improve upon it here. Thus the results shown in tables 1-2 should still be viewed as preliminary regarding the optical depth, which is essentially determined by the polarization data. Upcoming WMAP 2-year analysis/release of polarization data should elucidate the current situation. The code developed here will be made available to the community at \\texttt{cosmas.org}." }, "0403/astro-ph0403559_arXiv.txt": { "abstract": "We report the non-detection of the \\cii\\ \\fscii\\ 157.74 $\\mu$m transition in the $z=6.42$ quasar SDSS J1148+5251 after 37.5 hours of integration with the James Clerk Maxwell Telescope. This transition is the main cooling line of the star-forming interstellar medium, and usually the brightest FIR line in galaxies. Our observed 1 $\\sigma$ RMS = 1.3 mK in the $T_{A}^{*}$ scale translates to $\\Lcii<2.6 \\times 10^{9}$ L$_\\odot$. Using a recent estimate of the far-infrared continuum of this quasar, we derive for SDSS J1148+5251 $\\Lcii/\\Lfir<5\\times10^{-4}$, a ratio similar to that observed in local ultra-luminous infrared galaxies but considerably smaller than what is typical in nearby normal and starburst galaxies. This indicates that the small \\Lcii/\\Lfir\\ ratio observed locally in luminous far-infrared objects also persists at the highest redshifts. ", "introduction": "The most distant quasar in the universe known at the time of this writing is SDSS J114816.64+525150.3 (hereafter SDSS J1148+5251) at a redshift $z=6.42$ (Fan et al. 2003; Walter et al. 2003; Bertoldi et al. 2003a). With a bolometric luminosity of L$_{\\rm bol}\\sim10^{14}$ L$_\\odot$, this is an extremely luminous object powered by a $\\sim3\\times10^{9}$ M$_\\odot$ supermassive black hole at its core (Willott et al. 2003). The high far-infrared (FIR) luminosity of this object implies that the host galaxy is forming stars at the prodigious rate of 3000 M$_\\odot$ yr$^{-1}$ (Bertoldi et al. 2003b). At this redshift the universe is only 840 million years old and both the black hole mass and star formation rate imply the existence of a very massive galaxy formed from one of the rarest high-density peaks in the matter distribution. Although there are hints that SDSS J1148+5251 may be weakly gravitationally lensed (White et al. 2003), we assume no lensing magnification in this paper. The recent detection of bright molecular and FIR continuum emission from the host galaxy of SDSS J1148+5251 (Walter et al. 2003; Bertoldi et al. 2003a; Bertoldi et al. 2003b) prompted us to observe the \\cii\\ \\fscii\\ 157.74 $\\mu$m fine structure transition in this object. This is the main cooling transition of the star-forming interstellar medium (e.g., Tielens \\& Hollenbach 1985), and is commonly the single brightest emission line in galaxies. For example, Stacey et al. (1991) found that most galaxies emit 0.1\\% to 1\\% of their far-infrared luminosities in this line alone. Given its extremely high luminosity and its relationship to star formation activity, the redshifted \\cii\\ 158 $\\mu$m transition is attractive as a star formation indicator at high-$z$ (e.g., Stark 1997). This transition is conveniently placed into the atmospheric 1~mm window for redshifts $z\\gtrsim6.2$, making it accessible to ground-based instrumentation. Furthermore, a spectroscopic approach to detecting high-$z$ sources has the clear advantage of containing redshift information, unlike continuum observations, thus avoiding problems of source confusion that stem from the low angular resolution of single-dish radio telescopes equipped with bolometer arrays. As early as 1997, however, Infrared Space Observatory (ISO) observations suggested potential problems with this method. Indeed, Malhotra et al. (1997) and Luhman et al. (1998) found that the proportionality between the \\cii\\ and FIR continuum luminosities observed in nearby star forming galaxies broke down for luminous infrared galaxies, where the \\cii\\ luminosity appeared not to exceed $\\Lcii\\sim10^9$ \\Lsun. These conclusions, confirmed and expanded in a recent analysis of ISO data by Luhman et al. (2003), have since cast doubts on the usefulness of redshifted \\cii\\ to probe star formation in the distant universe. The observations presented here have bearing on this matter: does the paucity of \\cii\\ emission observed in local ultra-luminous infrared galaxies (ULIRGs) hold for objects in the early universe? ", "conclusions": "Our redshifted \\cii\\ observations at $\\nu\\sim256$~GHz have failed to detect the fine structure 158 $\\mu$m \\cii\\ transition in this high-$z$ quasar. We place a 1 $\\sigma$ limit on the \\cii\\ luminosity of SDSS J1148+5251 of $\\Lcii<2.6\\times10^9$ \\Lsun\\ (including 30\\% calibration uncertainty). Given the observed FIR flux of this source, this places an upper limit on its \\cii\\ to FIR ratio of $\\Lcii/\\Lfir<5\\times10^{-4}$, substantially lower than what is found in most local star-forming galaxies but similar to what is observed in nearby ultra-luminous IR galaxies. If the cause of this \\cii\\ deficit in SDSS J1148+5251 was a large FIR component due to partial dust reprocessing of the AGN radiation, thus unrelated to star formation activity, the $\\sim$3000 M$_\\odot$ yr$^{-1}$ star formation rate obtained from the FIR continuum could be a substantial overestimate. Obtaining a better SED for the FIR continuum emission, to constrain better the FIR luminosity, as well as improving the sensitivity of the CO \\jone\\ observations, may help to elucidate this matter. Will redshifted \\cii\\ observations open a new window onto high-$z$ galaxies? Our results are based on observations of just one object at high redshift. A larger sample of $z>6$ sources with high FIR luminosities must be observed in the \\cii\\ transition to establish whether low $\\Lcii/\\Lfir$ ratios are a general phenomenon. If a maximum luminosity $\\Lcii\\sim10^9$ \\Lsun\\ were confirmed for most sources, the Atacama Large Millimeter Array would still be able to detect the \\cii\\ line and spatially resolve its distribution and kinematics. The answer, then, is a hopeful {\\em yes}." }, "0403/astro-ph0403245_arXiv.txt": { "abstract": "{White dwarfs in globular clusters offer additional possibilities to determine distances and ages of globular clusters, provided their spectral types and masses are known. We therefore started a project to obtain spectra of white dwarfs in the globular clusters NGC~6397 and NGC~6752. All observed white dwarfs show hydrogen-rich spectra and are therefore classified as DA. Analysing the multi-colour photometry of the white dwarfs in NGC~6752 yields an average gravity of \\logg\\ = 7.84 and 0.53\\Msolar\\ as the most probable average mass for globular cluster white dwarfs. Using this average gravity we try to determine independent temperatures by fitting the white dwarf spectra. While the stellar parameters determined from spectroscopy and photometry usually agree within the mutual error bars, the low resolution and S/N of the spectra prevent us from setting constraints stronger than the ones derived from the photometry alone. For the same reasons the white dwarf spectra obtained for NGC~6397 unfortunately do not provide an independent distance estimate of sufficient accuracy to distinguish between the long and short distance scale for globular clusters. ", "introduction": "As white dwarfs are the final stage in the evolution of all low mass stars, many are expected to exist in globular clusters. However, due to their faintness and occurrence in very crowded fields they were detected only after the refurbishment of HST with the WFPC2. Several candidate white dwarfs were then soon identified in M15 (de Marchi \\& Paresce \\cite{depa95}), $\\omega$ Cen (Elson et al. \\cite{elgi95}), NGC 6397 (Paresce et al. \\cite{pade95a}, Cool et al. \\cite{copi96}), M4 (Richter et al. \\cite{rifa95}, \\cite{ribr02}), 47 Tuc (Paresce et al. \\cite{pade95b}, Zoccali et al. \\cite{zore01}) and NGC 6752 (Renzini et al. \\cite{rebr96}). Renzini \\& Fusi Pecci (\\cite{refu88}) suggested to use the white dwarf sequence as a standard candle for determining the distance to nearby globular clusters, similarly to the traditional main sequence fitting procedure using local subdwarfs with known trigonometric parallax. In this case, the white dwarf sequence of the cluster is compared to a sequence constructed with local white dwarfs with accurate trigonometric parallax. The method was then applied to NGC 6752 (Renzini et al. \\cite{rebr96}) and 47 Tuc (Zoccali et al. \\cite{zore01}, where also the result from Renzini et al. \\cite{rebr96} for NGC~6752 was updated). However, while the updated distance modulus for NGC~6752 agrees well with all other distance determinations for this cluster, the white dwarf distance to 47~Tuc is considerably shorter than that found by Gratton et al. (\\cite{grfu01}, \\cite{grbr03}) via main sequence fitting. While it may seem strange to use the faintest objects in a globular cluster to derive its distance white dwarfs offer some advantages as standard candles when compared to main sequence stars: \\begin{itemize} \\item They come in just two varieties - either hydrogen-rich (DA) or helium-rich (DB) -- {\\em independent of their original metallicity} and, in both cases, their atmospheres are virtually free of metals. So, unlike in the case of main sequence fitting, there is not the problem of finding local calibrators with the same metallicity as the globular clusters. \\item White dwarfs are locally much more numerous than metal-poor main sequence stars and thus allow to define a better reference sample. \\end{itemize} However, the method has its own specific problems, which are discussed in great detail in Zoccali et al. (\\cite{zore01}) and Salaris et al. (\\cite{saca01}). Indeed, the location of the white dwarf cooling sequence depends on: \\begin{itemize} \\item {\\em the white dwarf mass}\\\\ On theoretical grounds, given the observed maximum luminosity reached on the asymptotic giant branch (AGB), the mass of currently forming white dwarfs in globular clusters should be $0.53\\pm 0.02$\\Msolar\\ (Renzini \\& Fusi Pecci \\cite{refu88}, Renzini et al. \\cite{rebr96}). Unfortunately, there are no local white dwarfs in this mass range with directly determined masses (i.e. without using a mass-radius relationship). There is, however, a handful of local white dwarfs with spectroscopically determined masses near this value (cf. Table 1 in Zoccali et al. \\cite{zore01}), which allows to construct a semi-empirical cooling sequence for $M_{\\rm WD}$=0.53\\Msolar, once relatively small mass-dependent corrections are applied to each local white dwarf. The spectroscopic determination of the mass of white dwarfs in a globular cluster was first attempted by Moehler et al. (\\cite{mohe00}) for white dwarfs in NGC~6397. However, the low S/N of the spectra of these very faint stars did not allow to determine the mass with sufficient accuracy. \\item {\\em the white dwarf envelope mass}\\\\ In the case of DA white dwarfs the cooling sequence location depends also on the mass of the residual hydrogen-rich envelope. This affects {\\it spectroscopically} derived masses (see above), with the resulting mass being $\\approx 0.04$\\Msolar\\ higher when using the {\\it evolutionary} envelope mass ($\\approx 10^{-4}$M$_{\\rm WD}$, Fontaine \\& Wesemael \\cite{fowe97}) as opposed to virtually zero envelope mass. This mass uncertainty corresponds to an uncertainty of \\magpt{0}{1} in the distance modulus and 1--1.5 Gyr in the age derived from the main sequence turnoff. \\item {\\em spectral type}\\\\ DB stars are fainter than DA stars at a given colour, with the offset depending on the filter combination (i.e. the offset is greater in $V$ vs. $B-V$ than in $I$ vs. $V-I$). However, more massive DA white dwarfs are also fainter at a given colour. \\end{itemize} The white dwarf sequence also allows to determine the age of a globular cluster if its faint end is detected. In that case one can derive the age of the globular cluster from the luminosity of its oldest and faintest white dwarfs. Aside from the observational difficulties and the uncertainties in the cooling tracks (see Chabrier et al. \\cite{chbr00} for more details) any error in the assumed mass affects the result. Recent very deep HST observations allowed to detect the white dwarf cooling sequence in M~4 to unprecedented depths of $V$$\\approx$30 (Richer et al. \\cite{ribr02}). As a preliminary result Hansen et al. (\\cite{habr02}) derive an age of 12.7$\\pm$0.7 Gyr from the white dwarf luminosity function of M~4, consistent with other independent age estimates (their error bar, however, does not include errors due to the uncertainty of the white dwarf mass). Their result has recently been questioned by de Marchi et al. (\\cite{depa03}), who claim that the cluster membership of the white dwarfs can not be verified down to sufficiently faint limits to obtain more than a lower limit of the age. Richer et al. (\\cite{ribr04}) showed, however, that different methods of data reduction and analysis account for the different depths reached with the same data and thereby defended the original result of Hansen et al. (\\cite{habr02}). In view of the relevance of the white dwarf masses and spectral types to the problems described above we decided to observe spectra of white dwarfs in NGC~6397 and NGC~6752 in order to determine their spectral types and get mass estimates. Pilot observations of the white dwarf candidates in NGC~6397 (63.H-0348) had already shown that the targets are hydrogen-rich DA white dwarfs (Moehler et al. \\cite{mohe00}), but did not allow much quantitative work. ", "conclusions": "We observed a sample of white dwarfs in the globular clusters NGC~6397 and NGC~6752 and showed that they are all hydrogen-rich DA. From multicolour photometry we determined an average mass of 0.53$\\pm$0.03\\Msolar\\ (uncertainty due to uncertainties of 0.02 dex in the average \\logg\\ and of \\magpt{0}{05} in the distance modulus of NGC~6752). This value (based on the assumption of a thick hydrogen layer) is identical to that assumed by Renzini et al. (\\cite{rebr96}) on theoretical grounds. Therefore both observational and theoretical arguments strongly advocate against the use of 0.6~\\Msolar\\ (the mean mass of the local white dwarfs) or even more for the mass of hot white dwarfs in globular clusters when comparing observations to theoretical tracks. However, the limited S/N in combination with the low resolution of the spectroscopic data prevented the independent determination of masses from spectroscopic fits alone. Multi-colour photometry may be the better way to determine the physical parameters of white dwarfs in globular clusters once their spectral types are known. For spectroscopic observations our experience shows that crowding can severely limit the usefulness of the data due to problems with sky subtraction. We would therefore strongly recommend to look for white dwarf candidates in ground-based wide-field photometry to avoid the problems we encountered for NGC~6752. Also a better sensitivity in the blue would be useful for future spectroscopy. \\begin{table}[!h] \\caption[]{Results for the brightest object in NGC~6397 with \\logg\\ as free parameter. We also give M$_V$ from the model fit (assuming a thick hydrogen layer), the observed $V$ corrected for a reddening of E$_{\\rm B-V}$ = \\magpt{0}{18}, and the derived distance modulus. \\label{par-spec3}} \\begin{tabular}{lrrrrrr} \\hline \\hline & \\teff & $\\Delta$\\teff & \\logg & $\\Delta$ \\logg & M$_V$ & (m-M)$_0$ \\\\ & [K] & [K] & & & & \\\\ \\hline DK & 18800 & 340 & 7.72 & 0.06 & \\magpt{10}{47} & \\magpt{11}{72}\\\\ RN & 18700 & 520 & 8.09 & 0.12 & \\magpt{11}{02} & \\magpt{11}{17}\\\\ \\hline \\end{tabular} \\end{table}" }, "0403/hep-ph0403293_arXiv.txt": { "abstract": "We compute the neutrino detection rates to be expected at a low-energy beta-beam facility. We consider various nuclei as neutrino detectors and compare the case of a small versus large storage ring. ", "introduction": "The pioneering experiment of R. Davis \\cite{davis} has started the era of neutrino astronomy. Because they only have weak interactions with matter, neutrinos are precious messengers of what happens in the interior of stars, like our sun, or in explosive phenomena, such as Supernova type II explosions. Such astronomical neutrinos therefore provide an important source of information for our understanding of the life and death of stars. Nuclei are commonly used as detectors in neutrino observatories as well as in various experiments aiming at studying intrinsic neutrino properties, such as their masses and mixings. A precise knowledge of neutrino-nucleus cross-sections is needed for the interpretation of these measurements and/or to study the feasibility of new projects. The understanding of neutrino-nucleus interactions is also of crucial importance for various astrophysical processes. Timely examples include neutrino nucleosynthesis \\cite{haxton,woosley}, or the nucleosynthesis of heavy elements during the so-called r-process \\cite{bahanunucl,gail,bahaastro,qian,goriely}. If the latter takes place during the explosion of Supernovae type II, where a gigantic amount of energy is emitted as neutrinos of all flavors, final abundances depend on several nuclear properties, among which the interactions with neutrinos. According to existing simulations, the average energy of neutrinos emitted from core-collapse Supernovae is about $10$~MeV for electron neutrinos and about $20$~MeV for muon and tau neutrinos \\cite{raffelt}. Notice however that, due to oscillations, electron neutrinos can become hotter while traversing the star \\cite{hax1999,smirnov,nuPb}. The predicted spectra cover the $50$~MeV region and present a tail up to about $100$~MeV \\cite{raffelt}. Reactor and solar neutrinos have typical energies in the $10$~MeV energy range, while accelerator and atmospheric neutrinos cover the GeV and multi-GeV range. The various theoretical approaches employed to describe neutrino-nucleus interactions therefore involve nuclear as well as nucleonic degrees of freedom (for a review, see \\cite{revuekubo,jpg}). There are a number of open issues in this context. The A=2 system is the simplest case, for which the reaction cross sections can be estimated with high accuracy \\cite{kubodera}. However, there is still an important quantity, namely L$_{1,A}$, related to the axial two-body current, which dominates the theoretical uncertainty in neutrino-deuteron interactions. For heavier nuclei, in the tens of MeV energy range, the reaction cross sections are dominated by collective modes, like the Gamow-Teller resonance or the Isobaric Analog State, which have been extensively studied in the past \\cite{osterfeld}. As the neutrino impinging energy increases, transitions to states of higher multipolarity (such as the spin-dipole or higher forbidden transitions) become important \\cite{volpelead}. The latter also play an important role in the context of core-collapse Supernova physics \\cite{kolbe,gail,jon,volpelead}. Although some information on these states can be gathered through other probes, such as charge-exchange reactions \\cite{osterfeld}, muon capture \\cite{giai}, or inelastic electron scattering \\cite{electron}, the experimental information is rather scarce. Note that the understanding of neutrino-carbon reactions with neutrinos produced from the decay in flight of pions is still an open issue, for most of the theoretical calculations over-estimate the experimental value \\cite{c12}. So far, measurements with low-energy neutrinos have been performed in a few cases only, namely deuteron \\cite{deut}, carbon \\cite{expc12}, and iron \\cite{iron}. Systematic studies would be of great importance both for what concerns the interpolation from the MeV to the GeV neutrino energy range and the extrapolation to neutron-rich nuclei, as required in the astrophysical context. Neutrino-nucleus interaction studies were one of the main physics issues of the proposed ORLAND underground neutrino facility, which was based on a conventional neutrino source (pion and muon decays) \\cite{orland,jpg}. A smaller version of the ORLAND project is now under study \\cite{efremenko}. At present, the MINER$\\nu$A project \\cite{minerva} includes the study of neutrino-nucleus interactions for neutrino energies in the GeV range. Here, we study the potential of a low-energy neutrino facility based on beta-beams, a novel method to produce neutrino beams \\cite{zucchelli}. This consists in boosting exotic ions which decay through beta-decay and produce pure, collimated and well-understood electron neutrino fluxes. Such a method could be exploited for a future facility at CERN \\cite{zucchelli,mats}. High energy beta-beams would be fired to a gigantic Cherenkov detector like UNO \\cite{uno}, located in an (upgraded) Fr\\'ejus underground laboratory to study, in particular, the possible existence of CP violation in the leptonic sector \\cite{zucchelli,mats,mauro}. The discovery potential with a very high $\\gamma$ and a longer baseline is discussed in \\cite{jj,tmms}. It has recently been proposed to use the beta-beam concept for the production of low-energy neutrinos \\cite{lownu}. Several laboratories will produce intense exotic beams in the near future and could, therefore, be possible sites for a low-energy beta-beam facility. These include GANIL, CERN, GSI, as well as the EURISOL project. Low-energy neutrino beams would offer an interesting opportunity to study various neutrino properties, such as e.g.\\ the neutrino magnetic moment \\cite{munu}, as well as neutrino-nucleus interactions, of interest for nuclear physics, particle physics and astrophysics. In the former case, one would exploit the ions at rest as an intense neutrino source, whereas, in the latter case, one would use boosted ions, which would be collected in a storage ring \\cite{lownu}, as in the original high energy proposal. An important feature of such beta-beams is that the boost factor of the accelerated ions can be varied, allowing one to explore various neutrino energy ranges. In this paper, we present for the first time charged-current neutrino-nucleus interaction rates achievable at a low-energy beta-beam facility. We consider two possible cases for the dimensions of the storage ring, for which we inspire ourselves of the one planned in the future GSI facility \\cite{gsi} and the one thought in the CERN baseline scenario \\cite{zucchelli,mats}. We consider various target nuclei as neutrino detectors, namely deuteron, oxygen, iron and lead, which are commonly used in existing or planned experiments \\cite{orland}. Related work in the case of lead can be found in \\cite{gailnew}. ", "conclusions": "" }, "0403/hep-ph0403308_arXiv.txt": { "abstract": "We consider the infrared modification of gravity by ghost condensate. Naively, in this scenario one expects sizeable modification of gravity at distances of order 1000 km, provided that the characteristic time scale of the theory is of the order of the Hubble time. However, we argue that this is not the case. The main physical reason for the conspiracy is a simple fact that the Earth (and any other object in the Universe) has velocity of at least of order $10^{-3}c$ with respect to the rest frame of ghost condensate. Combined with strong retardation effects present in the ghost sector, this fact implies that no observable modification of the gravitational field of nearby objects occurs. Instead, the physical manifestation of ghost condensate is the presence of ``star tracks'' --- narrow regions of space with growing gravitational and ghost fields inside --- along the trajectory of any massive object. We briefly discuss the possibilities to observe these tracks. ", "introduction": "\\label{intro} Emerging evidence for the accelerated expansion of the Universe triggered interest in the non-standard theories of gravity in which gravitational interactions get modified in the infrared. To some extent the motivation is that in these theories, unlike in the case of pure cosmological constant, physics responsible for the cosmic acceleration may manifest itself in observations at smaller distance scales (e.g., in Lunar Ranging experiments \\cite{Lue:2002sw,Dvali:2002vf}). Naively, the most straightforward way to modify gravity at distance scale $r_c$ would be to give a graviton a mass $m_g\\sim r_c^{-1}$. However, conventional massive gravity suffers either from the presence of ghosts or from the loss of predictivity because of strong coupling at unacceptably low energy scales \\cite{Arkani-Hamed:2002sp}. Similar problems arise in multi-dimensional models, and it is not clear whether there exists a consistent {\\it quantum} brane world theory where gravity is modified in the infrared and predictive power is not lost at unacceptably low energy scale. Recently, an example of a theory which does not suffer from the above problems has been constructed \\cite{Arkani-Hamed:2003uy}. This theory, dubbed ``ghost condensate'', is somewhat similar to the Fierz--Pauli massive gravity. The difference is that the Fierz--Pauli mass term breaks reparametrization invariance completely, while in the ghost condensate theory only the time reparametrization invariance \\[ t\\to t+\\xi(t,x) \\] is broken, while the invariance under (possibly time-dependent) spatial diffeomorphisms is kept intact. As a result, the latter theory becomes formally reparametrization invariant after a single St\\\"uckelberg field is introduced, as opposed to four St\\\"uckelberg fields in the Fierz--Pauli gravity. The key difference between these two theories is that in the case of ghost condensate, decoupling limit exists in which gravity is switched off while the St\\\"uckelberg sector is still described by a well defined low-energy effective theory valid up to a certain energy scale $M$. The price one pays is that the Lorentz invariance is not preserved by ghost condensate; for instance, at the quadratic level one effectively adds to the Einstein theory the ``mass term'' of the form \\be \\label{mass} \\int dtd^3x{1\\over 8}M^4h_{00}^2\\;. \\ee As a consequence of this violation of the Lorentz invariance, the dispersion law for the St\\\"uckelberg field\\footnote{In what follows, we somewhat loosely refer to this field as a ghost. It is worth stressing however, that this is not a ghost field in the usual sense, i.e. the sign in front of its kinetic term in the action is positive.} $\\pi$ has rather peculiar form, \\[ \\omega^2={\\alpha\\over M^2} k^4\\;, \\] where $\\alpha$ is a dimensionless parameter of the theory. Another manifestation of the fact that the Lorentz invariance is broken is that modification of gravity in the infrared is not characterized by a single length scale $r_c$. Instead, there is a length scale $r_c$, equal to \\be \\label{rc} r_c={\\sqrt{2}M_{Pl}\\over M^2} \\ee and a time scale $t_c$ given by \\be \\label{tc} t_c={2M_{Pl}^2\\over \\alpha M^3}\\;. \\ee As discussed in Ref.~\\cite{Arkani-Hamed:2003uy}, for a massive source {\\it at rest} the length scale $r_c$ determines the characteristic distance at which the gravitational potential starts to deviate from the Newtonian one, while $t_c$ determines the characteristic time needed for this deviation to show up. Naively, this implies that, assuming that $\\alpha\\sim 1$ and $t_c$ is of the order of the present age of the Universe, $t_U\\sim 15$ Gyr, one might expect a sizeable modification of gravity due to ghost condensate at length scales of order 1000 km. However, it was noted already in Ref.~\\cite{Arkani-Hamed:2003uy} that the retardation effects are very strong in the ghost sector. The point is that it is the whole history of a system that determines its actual gravitational potential in the presence of ghost condensate. The purpose of this paper is to better understand this feature and thus reconsider possible observational signatures of ghost condensate. It is worth noting that for the moment, consequences of ghost condensate with $t_c \\sim t_U$ for present day cosmology have not been elaborated yet (see, however, Ref.~\\cite{Arkani-Hamed:2003uz} where ghost condensate was used to construct a model of inflation with quite unusual perturbation spectra). Still, we believe that the question we address is of relevance, since ghost condensate is an interesting infrared modification of gravity whose consistency is beyond any doubt. Surprisingly, we find that the above naive expectation is incorrect and it is not excluded that we live in the Universe with $t_c\\sim t_U$ and $r_c\\sim 1000$~km and have not noticed that so far. To understand how that could be, one recalls a simple fact that objects in the Universe are not at rest. Instead, solar system and other stars in our Galaxy rotate around the center of the Galaxy with typical velocity of order $10^{-3}$, while the Galaxy itself moves in the local cluster of galaxies with the velocity of the same order of magnitude. A well-known observational consequence of this motion is the dipole anisotropy of the cosmic microwave background. This implies that all stellar objects have velocity of at least of the same order of magnitude with respect to the rest frame of ghost condensate\\footnote{A priori one may think that this velocity can be significantly larger if the rest frame of the CMB has a finite velocity with respect to the rest frame of the ghost condensate. On the other hand we expect that one of the effects of the Hubble friction is to slow down this overall motion of the CMB. Throughout this paper we assume that the rest frame of the CMB coincides with the rest frame of the ghost condensate, leaving the study of whether it may be not the case for future.}. Now, it takes time of order $t_c$ for the modification of the gravitational potential to occur. Consequently, the effect of ghost condensate which we can observe on the Earth now is not a modification, say, of the gravitational field of the Sun, but the ``ghost'' tail of the potential of a star which was located nearby (in the rest frame of ghost condensate) time $t_c$ ago. The Universe with ghost condensate can be thought of as a kind of a bubble chamber where all moving massive objects leave long (and, as we will see later, narrow) tracks in which ghost field and gravitational potential are perturbed. The time delay between the moment when the object passes a given point in space and the appearance of the track around this point is of order $t_c$. This observation is our main result. The rest of this paper is organized as follows. In section \\ref{general} we derive a general expression for the gravitational potential of a massive source in the presence of ghost condensate (in the Newtonian approximation). In section \\ref{specific} we consider specific examples, namely a potential of a moving point-like source and the effect of the finite size of the source. Section \\ref{evidence} contains preliminary discussion of phenomenological implications of our calculations. Technical details can be found in Appendix. ", "conclusions": "\\label{evidence} Now we are at a point to discuss possible observational signatures of ghost condensate, taking into account the effect of non-zero velocity of all stellar objects in the rest frame of ghost condensate. Let us start with a few preliminary remarks. From the phenomenological point of view, ghost condensate (at the linearized level) has the characteristic time and length scales $t_c$ and $r_c$. These parameters are related to the two microscopic parameters, mass scale $M$ and dimensionless parameter $\\alpha$, as written in Eqs.~(\\ref{rc}), (\\ref{tc}). The allowed deviation of the latter parameter from unity is determined by the amount of fine-tuning that one is ready to introduce in the theory. We are going to be rather generous in this respect; in fact, our discussion is quite flexible and as large values of $\\alpha$ as $10^{10}$ will not affect it significantly. An exhaustive phenomenological analysis of the theory would have resulted in the exclusion plot in the $(t_c,r_c)$-parameter space, and in the detailed discussion of the characteristic observational signatures for different allowed regions. We believe that such an analysis deserves a separate publication. Our purpose here is to discuss qualitative features of the ghost condensate phenomenology, stressing the crucial role of the effect of finite velocity. Our claims are the following. \\begin{enumerate} \\item It is very unlikely to observe ghost condensate with $t_c\\sim t_U$, where $t_U$ is the present age of the Universe. In other words, it is crucial for the observability of ghost condensate that it enters the regime in which tracks with exponentially enhanced field are present. \\item The tracks of compact massive objects (stars) become pronounced earlier than the tracks of the supermassive objects of small density (galaxies). \\item The chance to observe ghost condensate is larger for larger values of $r_c$. \\item Relatively promising ways of searching for tracks in ghost condensate are: (i) search for ``mad'' stars which feel the gravitational field of the tracks of other stars; (ii) microlensing observations and (iii) gravitational wave experiments. \\item It may happen that $t_c$ is so small that tracks of some objects are already in the non-linear (quantum?) regime and, still, we have not noticed the presence of ghost condensate so far. Consequently, it may happen that to fully understand the phenomenology of ghost condensate one needs the details of the UV completion in the ghost sector. \\end{enumerate} Let us first explain why it is unlikely to observe ghost condensate for large characteristic time scales, $t_c\\gtrsim t_U$. In this regime, tracks with exponentially enhanced field did not have enough time to develop, so the only potentially observable effects are due to the gravitational potential in the wave zone. Let us first assume that the characteristic size $r_c$ is somewhat larger than the size of a typical star like the Sun, \\[ r_c\\gtrsim 10^6\\;\\mbox{km}\\; \\] This actually implies that the parameter $\\alpha$ is quite large, $\\alpha\\gtrsim 10^5$; however, as we will see, the chance to detect ghost condensate is even lower for smaller values of $r_c$. Then, to estimate the ``extra'' gravitational potential $\\Delta\\Phi$ (see, Eq.~(\\ref{NDel})) of a star whose trajectory was at a distance $y_0$ to the current location of the Earth (in the rest frame of ghost condensate) we make use of the expression (\\ref{weaklimit}) for the potential of a point-like source in the wave zone, \\be \\label{weakstar} \\Delta\\Phi\\sim 10^{-20} \\l \\mu\\over M_{\\odot}\\r \\l{ 10^{-3}\\over v}\\r \\l{r_c\\over y_0}\\r^2 \\sin{y_0\\Delta y\\over 2r_c^2} \\ee where we set $T=1$, expanded the phase of oscillations in Eq.~(\\ref{weaklimit}) near a given space-time point setting \\[ y=y_0+\\Delta y \\] neglected extremely slow variation of this phase in time and in $z$-direction and dropped the constant shift of this phase. Equation (\\ref{weakstar}) applies to the rest frame of ghost condensate, while in the rest frame of the Earth the gravitational field has the form of a wave of the amplitude $\\sim 10^{-20}$ and frequency \\be \\label{nu} \\nu= {v_r y_0\\over 4\\pi r_c^2}\\simeq 2\\cdot 10^{-5}\\; \\mbox{Hz}\\l {v_r\\over 10^{-3}}\\r \\l{y_0\\over r_c}\\r\\l{10^6\\;\\mbox{km}\\over r_c}\\r \\ee where $v_r$ is the Earth velocity in the direction transverse to the star trajectory. Note that this is a scalar wave unlike the tensor waves of the Einstein theory. The gravity waves of such a low frequency are in principle accessible to the LISA project (see, e.g., Refs.~\\cite{Thorne:1995xs,Lobo:2002pr} for reviews of the gravitational wave experiments), however, its sensitivity at these frequences is at the level\\footnote{It is worth noting that gravitational wave experiments are sensitive not to the amplitude $\\Delta\\Phi$ of the gravitational wave itself, but to the product $\\sqrt{n}\\Delta\\Phi$, where $n$ is a number of cycles produced in a logarithmic band about a given frequency. In the case at hand $n\\sim\\nu y_0/v_r=y_0^2/(4\\pi r_c^2)$. This comment is practically irrelevant for our discussion.} $\\Delta\\Phi\\sim 10^{-17}$. Furthemore, it is straightforward to see that the probability $p(r_c)$ to have even this weak signal is extremely low. Indeed, to estimate this probability, note first that the probability ${\\cal P}(r_c)$ that the distance from the current position of the Earth to the nearest trajectory of a star is smaller than $r_c$, is given by \\be \\label{probest} {\\cal P}(r_c)\\sim {N_{st}N_gvt_Ur_c^2\\over r_U^3}\\sim 10^{-15} \\ee where $N_{st}\\sim N_g\\sim 10^{11}$ are, respectively, the number of stars in a typical galaxy and the number of galaxies in the Hubble volume; $r_U\\sim 10^{28}$ cm, $v\\sim 10^{-3} $ and $r_c\\sim 10^6$~km. For long enough period of observation $t_o$, such that the Earth travels a distance $L_E$ much larger than $r_c$, the probability $p(r_c)$ is somewhat larger, \\[ p(r_c)\\sim {\\cal P}(r_c){L_E\\over r_c}\\sim 10^4\\cdot {\\cal P}(r_c)\\l {t_o \\over 1\\;\\mbox{yr} }\\r\\l{v\\over 10^{-3}}\\r\\l{10^6\\;\\mbox{km}\\over r_c}\\r \\] Since the amplitude in Eq.~(\\ref{weakstar}) rapidly decreases at $y_0 > r_c$, this expression determines the probability of having a non-negligible signal. Clearly, this probability is quite low. The sensitivity of gravitational wave detectors is higher in the higher frequency bands, so one might expect better signal for smaller $r_c$, and hence higher frequency $\\nu$. For instance, LISA will have sensitivity at the level $\\Delta\\Phi\\sim 10^{-20}$ in the frequency range $10^{-3}\\div 10^{-1}$~Hz. The waves of amplitudes and frequencies in this range would be generated for $y\\sim r_c\\sim 1000$~km. In this case, in order to avoid the suppression of the amplitude due to the effect of the finite size of the source (section \\ref{thicksub}), one should consider very compact sources like neutron stars. But the above estimate shows that we should be extremely lucky to have a trajectory of a neutron star at a distance of order 1000 km from the LISA facility. Let us now discuss the effects from objects of larger size, e.g. galaxies. One could expect two different types of signatures on these large distance scales. The first is the gravitational wave signals like those discussed above. The second is the modification of the gravitational dynamics at large scales due to extra contributions to the Newtonian potential. However, it is straightforward to see that the effect of averaging discussed in section \\ref{thicksub} kills all these signatures unless the value of the characteristic length scale $r_c$ (and, correspondingly, of the parameter $\\alpha$) is extremely large. Indeed, due to the large mass of a galaxy there is an extra factor of order $10^{12}$, as compared to a star, in the estimate (\\ref{weakstar}) for the gravitational potential. However, Eq.~(\\ref{wavebound}) tells us that there is an extra suppression by at least a factor \\be \\label{halofield} {r_c^2\\over L_g^2}\\sim 10^{-30}\\l{r_c\\over 1000\\;\\mbox{km}}\\r^2\\l{30\\;\\mbox{kpc} \\over L_g}\\r^2 \\ee where $L_g$ is the size of a galaxy. We see that the effects of ghost condensate are indeed very small for large objects of small density. The above arguments show that chance to detect ghost condensate is very low if the characteristic time scale $t_c$ is longer than, or equal to the present age of the Universe. This forces us to discuss the regime $t_c\\lesssim t_U$. This regime is rather dangerous, as the gravitational and ghost fields grow exponentially inside the tracks. Still, let us make a few general remarks, postponing the detailed discussion of the ghost condensate phenomenology in this regime for future. If the ratio $t_U/t_c$ is large enough, the tracks of galactic halos are very pronounced (say, the gravitational potential in the track is comparable to the typical gravitational potential between two interacting galaxies). The fraction of the Universe filled by these tracks is estimated as \\be \\label{halofraction} {\\Delta V_h\\over V_U}\\sim N_g{r^2_{halo}vt_u\\over t_u^3}\\sim 0.05 \\l{N_g\\over 10^{11}}\\r \\l{r_{halo}\\over 100\\mbox{ kpc}}\\r^2\\l{v\\over 10^{-3}}\\r \\ee where we estimated the size of a halo of a typical galaxy as 100 kpc. It is unlikely that this effect would have been unnoticed. For instance, the dynamics of a sizeable number of galaxies would have been affected by these tracks. For not so large $t_U/t_c$ the situation is more interesting. There is a range of parameters in which tracks of galaxies are unnoticeable, but star tracks are strong. As an example, one finds from Eqs.~(\\ref{weakstar}) and (\\ref{halofield}) (assuming $r_c=1000$~km for definiteness) that for \\[ {t_U\\over t_c}\\sim 2 \\ln{10^{20}}\\sim 92 \\] the gravitational potentials in the tracks of neutron stars are of order one, while the gravitational potentials in the tracks of typical galactic halos are still of order $10^{-18}$ which is more than 10 orders of magnitude smaller than the gravitational potential of a typical galaxy at a distance of order 1 Mpc. Consequently, there is an interesting range of the characteristic time scales, \\be \\label{interesting} 0.01 \\lesssim {t_c\\over t_U} <1 \\ee in which star tracks, but not galactic tracks, are pronounced (up to $\\Delta\\Phi\\sim 1$). Normally, the existence of exponentially growing field would mean that an extreme fine-tuning of the parameters is needed to have potentially observable effects without ruling out the theory completely. Clearly, to have $t_c$ in the range (\\ref{interesting}) one needs some fine-tuning, but not that strong as one might expect. We suggest here three potential signatures of the star tracks. First, one can search for ``mad'' stars, intersecting the tracks of other stars, so that their motion is strongly affected by the gravitational field of the track (probably, just for a short period of time). Second, it seems possible to observe tracks in the microlensing experiments (see, e.g., Ref.~\\cite{Roulet:1996ur} for a review of microlensing experiments) due to the variation of the visible luminosity of a background star when the line of sight crosses the track. Finally, the gravitational wave detetors may detect a signal from the track, as discussed above. To observe mad stars one needs a galaxy whose disc is intersecting a track of the disc of another galaxy at the moment of observation. Plugging the typical disc size $r_{disc}\\sim 10$~kpc instead of $r_{halo}$ into Eq.~(\\ref{halofraction}) we find that this happens for approximately one galaxy of a thousand. To get a feeling of numbers let us also estimate the fraction of the volume of such a galaxy filled by the star tracks, \\be \\label{starfraction} {\\Delta V_{s}\\over V_{galaxy}}\\sim N_s\\l {r_s\\over r_{disc}}\\r^2\\sim 10^{-12}\\;, \\ee where $r_s\\sim 10^6$~km is the radius of a typical star. This estimate is not very optimistic as it implies that there is about one mad star in the galaxy at each moment of time. Note however, that one can significantly enhance the success probability by performing monitoring of stars for a long period of time. Also it would be interesting to check whether trapping of a star by the gravitational field of a track is possible. Finally, the estimate (\\ref{starfraction}) assumes that the parameter $r_c$ is smaller than $r_s$ so that the diameter of the star track is equal to the size of the star. At larger values of $r_c$ this fraction is enhanced by a factor of $(r_c/ r_s)^2$. Similar considerations show that chance to observe two other signatures (microlensing and gravity waves) are rather low for $r_c$ in the range 0.3--0.5; see Fig.~\\ref{fig:de-gamma}). At higher luminosities, core galaxies can appear to have $\\gamma > 0.3$ if the core is not adequately resolved (either due to distance or to inner truncation of the profile by, e.g., dust). (We \\textit{do} classify two galaxies in our sample as ``possible core'' galaxies, but these are clearly cases of inadequate resolution.) Although we have not yet attempted to model the complete profiles of \\textit{bulges}, it is reasonable to extend our results to them. \\citet{balcells03} have already done this for a sample of early-type bulges in the near-IR, using NICMOS data in conjunction with ground-based imaging. They find that the complete bulge profiles, after accounting for the presence of the outer disk, can be well modeled by S\\'ersic profiles, plus optional nuclear components (corresponding to, e.g., nuclear star clusters or point sources). This is in excellent agreement with our hypothesis that the profiles of lower-luminosity ellipticals and bulges are fundamentally S\\'ersic profiles, and promises to resolve a number of ambiguities and ``dichotomies'' reported in the literature. For example, \\citet{carollo97} and \\citet{seigar02} argue for a dichotomy between $R^{1/4}$ and exponential bulges, with the latter having low $\\gamma$ in contrast to the high $\\gamma$ of $R^{1/4}$ bulges and moderate-luminosity ellipticals. This is naturally explained if most bulges actually have S\\'ersic profiles (as is well supported by a number of studies) \\textit{and} if these S\\'ersic profiles extend into the nuclear region. The division between $R^{1/4}$ (S\\'ersic index $n = 4$) and exponential ($n = 1$) bulges is probably an artificial one, given that bulges in reality show a range of values of $n$. But as \\nocite{paper1}Paper~I shows, bulges with larger $n$ will have higher values of $\\gamma$ than bulges with low $n$. Thus, ``$R^{1/4}$'' bulges (higher $n$) will exhibit larger values of $\\gamma$ than ``exponential'' (lower $n$) bulges. Since bulge $n$ decreases along the Hubble sequence, the trend of decreasing $\\gamma$ with Hubble type noted by \\citet[][ their Fig.~3]{seigar02} follows as well. In retrospect, we can see that most of the early \\textit{HST} studies of galaxy centers, and some of the more recent ones \\citep[e.g.,][]{rest01,ravindranath01}, have focused on relatively high-luminosity systems. These samples thus included a mix of S\\'ersic galaxies with high $n$ values and genuine core galaxies, making a distinction between core and ``power-law'' galaxies based purely on $\\gamma$ feasible. More recent studies aimed at low-luminosity systems \\citep[e.g.,][]{carollo98,stiavelli01,seigar02} have since uncovered evidence for the low-$n$--low-$\\gamma$, high-$n$--high-$\\gamma$ trend that pure S\\'ersic profiles generate, and thus show that discriminating core galaxies purely by $\\gamma$ is problematic at best. \\subsection{Core Identifications and Core Parameters} %[x] \\label{sec:core-params} We find that most of the previously identified ``core'' galaxies in our sample \\textit{do} have distinct cores with shallow, power-law cusps. These cores stand out as downward deviations from the outer S\\'ersic profiles. Fitting with the \\bomba{} model provides a more natural, less ambiguous definition for ``true'' cores, without the possibility of misclassifying low-$n$ S\\'ersic profiles as cores. We are also able to re-classify one of the ``intermediate'' galaxies (NGC~5557) of \\citet{rest01} as a core galaxy. The two ambiguous galaxies --- NGC~3613 and NGC~5077 --- are simply cases where the apparent break radius is very close to the inner limits of the data. For NGC~3613, this is because the apparent core is close to the resolution limit (in fact, $r_{b}$ from the \\bomba{} fits is $< 0\\farcs16$ and thus smaller than the suggested resolution-based limit of \\nocite{faber97}Faber et al.\\ 1997). For NGC~5077, on the other hand, \\citet{rest01} clipped their data at $r = 0\\farcs1$ because of an apparent nuclear excess at smaller radii. A future fit including data at smaller radii and using an extra nuclear component to account for this excess, may help determine if NGC~5077 truly possesses a core. While our overall agreement with the core/non-core classifications of \\citet{lauer95} and \\citet{rest01} is quite good for the galaxies we analyze, \\textit{we find that Nuker-law fits systematically overestimate the size of the cores}: our break radii are $\\sim 1.5$--4.5 times smaller in size than the break radii from the published Nuker-law fits. Consequently, $\\mu_{b}$ values are brighter as well. We also find consistently higher values of $\\gamma$, though the difference is not as dramatic (see Table~\\ref{tab:core-params} and Figure~\\ref{fig:core-core}). This is in excellent agreement with the arguments of \\nocite{paper1,paper3}Papers~I and III: \\textit{all} Nuker-law parameters are sensitive to the radial size of the region where the fit is made. All parameters of the Nuker model, including $\\gamma$ and $r_{b}$, must be adjusted in order to fit both the core \\textit{and} the (non--power-law) part of the profile outside, with its intrinsic (S\\'ersic) curvature. Table~\\ref{tab:core-params} shows that, on average, the \\bomba{} values of $\\gamma$ match the \\textit{observed} core slope $\\gammap$ (as determined by \\citet{rest01}) better than the Nuker-law values do. The currently favored theory for core formation is the ejection of core stars by 3-body encounters with a decaying black hole binary formed following a merger of two galaxies with central supermassive black holes. Various calculations \\citep{ebisuzaki91,quinlan97,milosavljevic01} have estimated the stellar mass ejected during this process (\\mej), and generally find it to be $\\sim \\mbh$, where \\mbh{} is the mass of the resulting central black hole formed by the (assumed) coalescence of the binary. However, attempts to test these predictions by estimating \\mej{} from observed cores and comparing it with various estimates of \\mbh{} consistently produce values of $\\mej > \\mbh$. \\citet{faber97} found $\\mej = 3.5$--6.4 \\mbh; using more accurate estimates of \\mbh, \\citet{milosavljevic01} found $\\mej \\approx 1$--20 \\mbh. \\citet{ravindranath02} used the prescription for \\mej{} of Milosavljevi\\'{c} \\& Merritt and a much larger data set; they found $\\mej \\approx 2$--20 \\mbh{} at the low-mass end ($\\mbh \\sim 10^{8} M_{\\sun}$), while at the high-mass end ($\\mbh \\sim 10^{9} M_{\\sun}$) $\\mej \\approx 6$--25 \\mbh. Even considering only the galaxies with \\textit{measured} \\mbh, $\\mej/\\mbh \\approx$ 4--13. Milosavljevi\\'{c} \\& Merritt pointed out that the \\textit{total} ejected mass should increase with the number of mergers, but the observed ratios still seem high, particularly at the low-mass end, where there have presumably been fewer mergers. All of the studies cited above used parameters from Nuker-law fits to estimate \\mej. Since the estimated \\mej{} scales with $r_{b}$ --- in the parameterization introduced by \\citet{milosavljevic01} and used by Ravindranath et al.\\ (2002), $\\mej \\propto r_{b}$ --- overestimating $r_{b}$ will naturally overestimate \\mej. Thus at least some of the discrepancy between observed and predicted $\\mej/\\mbh$ is probably due to the tendency of Nuker-law fits to overestimate $r_{b}$, as we have found. Assuming that the core radii from \\bomba{} fits are typically $\\sim 2$--4 times smaller than the Nuker-law values, as is the case for our sample, $\\mej/\\mbh$ values should go down by comparable factors, which would put them in better agreement with the theoretical predictions. One of our core galaxies (NGC~4291) was noted by \\citet{ravindranath01} for possibly having an isothermal core (with $\\gamma = 0$), on the basis of their Nuker-law fits to a NICMOS image. The Nuker-law fit in \\citet{rest01} to the WFPC2 profile also has $\\gamma = 0.0$, which might seem to strengthen the case for an isothermal core. However, we find $\\gamma = 0.14$ from our \\bomba{} fit, which agrees very well with $\\gammap = 0.13$ determined by Rest et al. So the core of NGC~4291 is probably \\textit{not} isothermal. In Figure~\\ref{fig:core-global} we show the relation between the core and the global properties of the galaxies in our sample. We also indicate the upper limits on possible core radii for the S\\'ersic galaxies, based on the radii of the innermost valid data. For those galaxies where a clear core has been measured, we find that the relation between the break radius and the effective radius is approximately given by $r_b = 0.014 r_e$. This is a factor of two smaller than the relation found by \\citet{faber97}, consistent with our finding that fitting with the Nuker law tends to overestimate core sizes. There is a suggestion of a weak trend of $r_{b}$ increasing with galaxy luminosity, which would be in agreement with what Faber et al.\\ found \\citep[see also][]{laine03}, but for our sample this ``trend'' is anchored by only two points, so it is dubious. Unfortunately, the narrow magnitude range spanned by the core galaxies in our sample ($\\lesssim 1.5$ mag) precludes a proper test of the magnitude-$r_{b}$ relation reported Faber et al., which is based on galaxies spanning $\\gtrsim 3$ mag (and the composite trend in Fig.~9 of Laine et al.\\ spans almost 5 magnitudes). There is no clear magnitude-related trend in the \\textit{ratio} of our $r_{b}$ measurements to the Nuker-law measurements, which suggests that the magnitude-$r_{b}$ trend may be unaffected by changes in $r_{b}$, except possibly in the scatter. However, a proper evaluation of how the magnitude-$r_{b}$ relation is affected by better measurements of $r_{b}$ must await \\bomba{} fits to a larger sample of core galaxies. There is \\textit{no} evidence for a relationship between $n$ and $r_{b}$; this may be partly due to large uncertainties in $n$ \\nocite{caon93}(Caon et al.\\ 1993 found typical errors of $\\sim 25$\\% when fitting S\\'ersic profiles). Finally, we find \\textit{no} clear correlation between $\\gamma$ and the global properties of the core galaxies analyzed. This is agreement with what previous studies have found for core galaxies (e.g., \\nocite{rest01}Rest et al.\\ 2001, Figure~7; \\nocite{ravindranath01}Ravindranath et al.\\ 2001, Figure~3; \\nocite{laine03}Laine et al.\\ 2003, Figure~6; and the core galaxes in Figure~\\ref{fig:gamma-core} of this paper). \\subsection{Hidden Cores and the Core-Galaxy Fraction} %[x] An interesting point is to consider how well-resolved the underlying profiles of the various galaxies actually are. In several cases, \\citet{byun96} and \\citet{rest01} excluded points at small radii from their fits, usually due to the presence of significant nuclear dust or a distinct nuclear component (e.g., a nuclear point source). Thus, not all of the profiles take full advantage of \\textit{HST} resolution. While the nuclear components may include cases of nuclear star clusters, which make discussions of the underlying stellar profile ambiguous, the presence of dust means that some ``power-law'' (i.e., S\\'ersic-profile) galaxies could have hidden cores. If we divide the sample into two groups --- galaxies where the innermost valid data point is at $r < 15$ pc (spatially well resolved centers); and galaxies where the innermost valid point is at $r > 15$ pc (less well-resolved centers) --- we find that the \\textit{less} resolved galaxies are almost all\\footnote{The exceptions are NGC~4168 (core) and NGC~5077 (possible core).} well fit using just the S\\'ersic model. This suggests that at least some of the S\\'ersic galaxies could have ``hidden'' cores. This is not a new argument, obviously, as many authors have pointed out that ``power-law'' galaxies could include unresolved cores --- but it is interesting to consider how \\textit{few} of the S\\'ersic galaxies in our sample can really be declared free of \\textit{HST}-resolvable cores. Of the 21 galaxies, seven clearly have cores, two have possible cores (NGC~3613 and NGC~5077, see Section~\\ref{sec:core-params}), and only five (NGC~4478, NGC~5576, NGC~5796, NGC~5831, and NGC~5845) are clearly free of significant ($r_{b} > 5$ pc) cores. So in the limited range of absolute magnitude spanned by our full sample ($-18.3 \\gtrsim M_{B} \\gtrsim -21.4$), 33\\% of the galaxies have unambiguous, \\textit{HST}-resolved cores; but this is clearly a lower limit. The core fraction rises to 43\\% if we include the two possible cases, and in principle could be as high as 76\\%. It is also interesting to note that we can see in the absolute magnitudes a hint of the well-known dichotomy between core and non-core galaxies \\citep[see, e.g., the discussion in ][]{rest01}, even in our limited sample. This can be seen in Figure~\\ref{fig:core-global}, where the five \\textit{fully resolved} S\\'ersic galaxies tend to be fainter than the core galaxies; a Kolmogorov-Smirnov test gives a 95\\% probability that the two groups of galaxies come from different parent luminosity distributions. We have successfully fit the complete surface-brightness profiles of 19 out of 21 elliptical galaxies, from the \\textit{HST}-resolved central regions ($r \\sim 0\\farcs02$) out to $\\sim$ twice the half--light radius, using either: a) a pure S\\'ersic profile; or b) a ``\\bomba{}'' model consisting of an outer S\\'ersic profile joined to an inner power-law core. The former fits correspond to so-called ``power-law'' galaxies, which are perhaps better described as ``S\\'ersic galaxies,'' and the latter correspond to core galaxies. The combined use of these two models lets us address the following questions: \\begin{enumerate} \\item \\textit{How can we relate the central, HST-resolved part of the galaxies' surface-brightness profiles to the outer regions?} We show that most power-law ellipticals are well described at all radii by the simple S\\'ersic law (modulo any nuclear disks, etc.). On the other hand, core galaxies are extremely well fit with the \\bomba{} model. We find little need for a significant transition region between the outer (S\\'ersic) part of the \\bomba{} profile and the (power-law) core; any such transition region is small compared to the size of the core. \\item \\textit{Is there a dichotomy in nuclear profiles between low- and high-luminosity bulges and ellipticals?} Some recent \\textit{HST} studies have suggested that the apparent trend seen in intermediate- and high-luminosity bulges and ellipticals --- cores with shallow logarithmic slopes in high-luminosity systems, steeper nuclear slopes in lower-luminosity (``power-law'') systems --- breaks down at lower luminosities, because fainter bulges and dwarf ellipticals have shallow nuclear slopes. We show that the power-law galaxies in our sample have S\\'ersic profiles that extend into the limits of \\textit{HST} resolution, with $n \\sim 4$--6; this naturally explains the steep nuclear slopes previously reported. When combined with the well-known correlation between $n$ and luminosity, we can see that \\citep[as argued by][]{graham-guzman03} the general trend is most likely one of pure S\\'ersic profiles (plus possible extra components such as nuclear star clusters and disks), extending from low-luminosity systems with low-$n$ S\\'ersic profiles --- and thus shallow nuclear slopes --- to high-luminosity systems with high-$n$ profiles and steeper nuclear slopes. Only the high-luminosity \\textit{core} galaxies break the trend, due to the existence of the cores themselves. \\item \\textit{How can we unambiguously identify cores in galaxy profiles?} As we demonstrate, the traditional definition of cores using parameters from Nuker-law fits to galaxy profiles ($r_b \\geq 0\\farcs16$ and $\\gamma < 0.3$) leads to the real possibility of misclassifying galaxies with sufficiently shallow slopes (for example, exponential profiles) as core galaxies. \\textit{We define core galaxies as those possessing a well-resolved downward deviation from the inward extrapolation of the outer (S\\'ersic) profile}. This definition recovers previous core definitions for the high-luminosity ellipticals in our sample, but is immune to the danger of identifying exponential-like profiles as having cores. \\item \\textit{How can we more accurately determine the structural properties of cores?} As demonstrated in Paper~I, the Nuker law requires a broad, smooth transition (low values of $\\alpha$) between its two power-law regimes in order to fit the inner profiles of core and power-law galaxies, because this is the only way to reproduce the observed curvature of actual galaxy profiles. We find that this causes the core-size measurements (i.e., the break radius) to be overestimated by factors of 1.5--4.5 in comparison to the values derived by using the \\bomba{} model, which directly accounts for the intrinsic curvature of galaxy profiles. We also find that the logarithmic slope $\\gamma$ of the observed core is more accurately recovered with the \\bomba{} model. Using the smaller values we find, especially for $r_{b}$, should bring estimates of the ejected stellar mass due to core formation more in line with theoretical predictions. \\end{enumerate}" }, "0403/astro-ph0403251_arXiv.txt": { "abstract": "A soft component of thermal emission is very commonly observed from the surfaces of quiescent, accreting neutron stars. We searched with {\\it Chandra} for such a surface component of emission from the dynamical black-hole candidate XTE~J1118+480 (=\\ J1118), which has a primary mass $M_{1} \\approx$ 8 \\msun. None was found, as one would expect if the compact X-ray source is a bona fide black hole that possesses an event horizon. The spectrum of J1118 is well-fitted by a simple power-law model that implies an unabsorbed luminosity of $L_{\\rm x} \\approx 3.5 \\times 10^{30}$ \\lum~(0.3--7 keV). In our search for a thermal component, we fitted our {\\it Chandra} data to a power-law model (with slope and $N_{\\rm H}$ fixed) plus a series of nine hydrogen-atmosphere models with radii ranging from 9/8 to 2.8 Schwarzschild radii. For the more compact models, we included the important effect of self-irradiation of the atmosphere. Because of the remarkably low column density to J1118, $N_{\\rm H} \\approx 1.2 \\times 10^{20}$~cm$^{-2}$, we obtained very strong limits on a hypothetical thermal source: $k\\Tinf <$ 0.011 keV and $L_{\\infty,\\rm th} < 9.4 \\times 10^{30}$ \\lum~ (99\\% confidence level). In analogy with neutron stars, there are two possible sources of thermal radiation from a hypothetical surface of J1118: deep crustal heating and accretion. The former mechanism predicts a thermal luminosity that exceeds the above luminosity limit by a factor of $\\gtrsim 25$, which implies that either one must resort to contrived models or, as we favor, J1118 is a true black hole with an event horizon. In addition to neutron stars, we also consider emission from several exotic models of compact stars that have been proposed as alternatives to black holes. As we have shown previously, accreting black holes in quiescent X-ray binaries are very much fainter than neutron stars. One potential explanation for this difference is the larger and hence cooler surface of an 8 \\msun~compact object that might be masked by the ISM. However, our upper limit on the {\\it total} luminosity of J1118 of $1.3 \\times 10^{31}$ \\lum~is far below the luminosities observed for neutron stars. This result strengthens our long-held position that black holes are faint relative to neutron stars because they possess an event horizon. ", "introduction": "In principle it is possible to detect the radiation emitted from the surface of any ordinary astronomical body such as a planet or a star of any kind. On the other hand, it is quite impossible to detect any radiation from an event horizon, which is the immaterial surface of infinite redshift that defines a black hole. This is unfortunate because demonstrating the reality of the event horizon is a problem central to physics and astrophysics. Nevertheless, despite the complete absence of any emitted radiation, it is possible to marshal strong circumstantial evidence for the reality of the event horizon. One fruitful approach is based on comparing low mass X-ray binaries (LMXBs) that contain black hole primaries with very similar LMXBs that contain neutron star primaries. In the quiescent state of these systems (McClintock \\& Remillard 2004), the lack of a stellar surface leads to predictable consequences, such as the faintness of black holes relative to neutron stars (Narayan, Garcia, \\& McClintock 1997b; Garcia et al. 2001; Narayan, Garcia \\& McClintock 2002), and also the lack of a thermal component of emission from black holes, which is commonly present in neutron stars (this work). Similarly, in the outburst state, the presence of a surface in the neutron star (NS) systems gives rise to some distinctive phenomena that are absent in the black hole (BH) systems: (1) type I thermonuclear bursts (Narayan \\& Heyl 2002); (2) high-frequency ($\\sim1$~kHz) timing noise (Sunyaev \\& Revnivtsev 2000); and (3) a distinctive spectral component from a boundary layer at the stellar surface (Done \\& Gierlinski 2003). In quiescence, almost all accreting neutron stars (e.g., Cen X--4, Aql X--1, and KS 1731--260) display a soft (kT $\\sim$ 0.1 keV) thermal component of emission (see \\S4.2). The source of the thermal energy is uncertain; for example, it may be energy liberated by the impact of accreting matter (Narayan et al. 1997b, 2002) or crustal energy from the star's interior (Brown, Bildsten, \\& Rutledge 1998; Rutledge et al. 2002b). In any case, however, the observed X-ray luminosities, temperatures and distances of these NSs indicate that one is observing thermal emission from a source of radius $\\sim10$~km, which is plainly the stellar surface. On the other hand, no quiescent thermal component of emission has been reported for any of the 15 BH LMXBs (McClintock \\& Remillard 2004), which is the expected result if they possess an event horizon. These BH spectra are well-represented by a simple power law with photon index $1.5 < \\Gamma < 2.1$ (McClintock \\& Remillard 2004). However, a soft component of emission might have escaped detection for several reasons. For example, the quiescent BH LMXBs are fainter, making a soft component more difficult to observe. Also, compared to the surface of a NS, a hypothetical material surface surrounding a $\\sim10$~\\msun~compact object would be larger and therefore have a correspondingly lower surface temperature. Herein we search for a thermal component of emission in the quiescent spectrum of XTE J1118+480 (hereafter J1118), a BH LMXB with an extraordinarily high Galactic latitude ($b = 62$\\deg) and correspondingly low interstellar absorption: $N_{\\rm H}~\\approx 1.2~\\times~10^{20}$~cm$^{-2}$ (see \\S2). For this nominal column depth, the transmission of the interstellar medium (ISM) is 70\\% for the softest X-rays (0.3 keV) that we consider. Thus, J1118 provides a unique opportunity to search for a soft thermal component of X-ray emission. For the purposes of this study, we adopt a mass for the BH primary of $M_1 = 8$ \\msun~(McClintock et al. 2001a; Wagner et al. 2001; Orosz et al. 2004). The quiescent X-ray luminosity is $\\approx3.5 \\times 10^{30}$ \\lum~ (0.3--7 keV; D = 1.8 kpc), which is $10^{-8.5}$ of the Eddington luminosity (McClintock et al. 2003). Both the luminosity and the photon spectral index, $\\Gamma~= 2.02~\\pm~0.16$, are typical for a quiescent BH LMXB with a short orbital period, $P_{\\rm orb} = 4.1$~hr (McClintock et al. 2003). All of the BH and NS LMXBs that we consider herein are X-ray novae (a.k.a. soft X-ray transients) that undergo bright outbursts lasting several months, which are followed by years or decades of quiescence. During its outburst maximum, J1118 was exceptionally underluminous in the 2--12 keV band, $L_{\\rm x} \\approx 3 \\times 10^{35}$ \\lum, compared to the other BH LMXBs in outburst, $L_{\\rm x} \\sim 10^{38}$ \\lum (McClintock \\& Remillard 2004). In this work, we determine a strong upper limit on any soft thermal component in the {\\it Chandra} X-ray spectrum of J1118. Using our upper limit on this emission component, and assuming that J1118 possesses a hypothetical material surface, we set stringent temperature and luminosity upper limits on thermal emission from this surface for a wide range of assumed surface radii; we compare these limits to the observed temperatures and luminosities of quiescent NS LMXBs. We conclude that the absence of a soft thermal component of emission in the spectrum of J1118 rules strongly against the presence of a material surface and hence argues for the existence of an event horizon. This work is organized as follows. In \\S2 we examine the central question of the column density to J1118, and in \\S3 we discuss the observations, data analysis and model-fitting techniques. The development and computation of stellar atmospheric models appropriate to a compact and massive star are presented in Appendix A; the models include the effects of self-irradiation for stars so compact as to lie within their own photon spheres. In \\S4, upper limits on the temperature and luminosity of a thermal component from J1118 are summarized and discussed, and these results are compared to the thermal spectra observed for neutron stars. In \\S5, we interpret the absence of thermal emission form J1118 in terms of two conventional sources of thermal emission from NSs; in addition we consider emission from exotic models of massive, compact stars that have been proposed as alternatives to BHs. Our conclusions are summarized in \\S6. J1118 and the other BHs referred to throughout this work are among the 18 dynamically-confirmed BHs; for a review of the properties of these massive, compact X-ray sources, see McClintock \\& Remillard (2004). ", "conclusions": "We have examined the possibility that the dynamical black-hole candidate J1118 possesses a material surface rather than an event horizon. Either accretion onto such a surface or deep crustal heating would be expected to produce a quiescent thermal component of emission like those commonly observed for neutron stars. We have fitted our Chandra spectrum of J1118 to a model consisting of a fixed power-law component plus an atmospheric thermal component with variable temperature, $\\Tinf$. The spectral fits were repeated for a series of nine atmospheric models with radii ranging from the minimum allowable, $9/8\\, \\Rs$, to a maximum of $2.8\\, \\Rs$. For the most compact of these models, which lie within their own photon spheres, the self-irradiation of the atmospheres was taken into account. No emission in excess of a simple power-law component was detected in J1118, and very strong upper limits were set on the presence of a thermal source: $k\\Tinf <$ 0.011 keV and $L_{\\infty,\\rm th} < 9.4 \\times 10^{30}$ \\lum~(99\\% confidence level). If one assumes that the hypothetical crust of J1118 is composed of normal nuclear matter, then this stringent limit on a thermal component of luminosity is hard to reconcile with the theory of deep crustal heating and the observed fluence of J1118 during its outburst in 2000: The predicted quiescent luminosity exceeds the above limit on $L_{\\infty,\\rm th}$ by a factor of $\\gtrsim 25$. Possibly a contrived model of deep crustal heating and/or an extreme model of neutrino cooling of the core could explain this difference. On the other hand, if J1118 possesses a material surface and accretion powers the thermal emission seen from NSs, then one expects J1118 to have a luminosity at least as great as that of an average NS, whereas its {\\it total} luminosity in Eddington-scaled units is about 100 times less than the luminosity of a typical NS and fully 10 times less than the luminosity of even SAX J1808.4-3658 (Fig. 4). The above limit on thermal emission, in combination with the observed power-law emission, yields a very tight limit on the {\\it total} quiescent luminosity of $1.3 \\times 10^{31}$ \\lum, which is far below the luminosities observed for NSs. Because of the high transparency of the ISM, our results rule out the possibility that the total luminosity of J1118 could be augmented significantly by any ultrasoft component of emission. Thus J1118 -- and by inference the other dynamical BH candidates -- are truly faint relative to NSs (Figs. \\ref{fig:4} and \\ref{fig:5}). In summary, a sensitive search has failed to detect any thermal emission from a hypothetical surface surrounding J1118, although NSs very commonly show such surface emission due to either deep crustal heating or accretion. Our sensitivity to a thermal component of emission from J1118 is much greater than the emission predicted by the theory of deep crustal heating, assuming that J1118 has a material surface analogous to that of NSs. Likewise, there is no evidence that accretion is occurring in quiescence onto the surface of J1118, which is the mechanism often invoked to explain the far greater thermal luminosities of NSs. The simplest explanation for the absence of any thermal emission is that J1118 lacks a material surface and possesses an event horizon. Finally, our limits on thermal emission from J1118 rule out the possibility that there is a heretofore unseen and appreciable soft component of luminosity. This result implies that the dynamical BH candidates are truly faint relative to NSs and underscores our original argument that these compact objects have event horizons and are therefore genuine black holes (Narayan et al. 1997b). As discussed in \\S 5.3, however, we cannot at this time rule out certain very exotic alternatives. \\appendix" }, "0403/astro-ph0403298_arXiv.txt": { "abstract": "The masses of central massive black holes in BL Lac objects are estimated from their host galaxy absolute magnitude $M_{\\rm R}$ at $R$-band by using the empirical relation between absolute magnitude of host galaxy $M_{\\rm R}$ and black hole mass $M_{\\rm bh}$. Only a small fraction of BL Lac objects exhibit weak broad-line emission, and we derive the sizes of the broad-line regions (BLRs) in these BL Lac objects from the widths of their broad emission lines on the assumption of the clouds being virilized in BLRs. It is found that the sizes of the BLRs in these sources are usually 2-3 orders of magnitude larger than that expected by the empirical correlation $R_{\\rm BLR}-\\lambda L_{\\lambda}$(3000\\AA) defined by a sample of Seyfert galaxies and quasars \\citep{mj02}. We discuss a variety of possibilities and suggest it may probably be attributed to anisotropic motion of the BLR clouds in these BL Lac objects. If the BLR geometry of these sources is disk-like, the viewing angles between the axis and the line of sight are in the range of $\\sim 2^\\circ-12^\\circ$, which is consistent with the unification schemes. ", "introduction": "There is a tight correlation between the black hole mass $M_{\\rm bh}$ and the stellar dispersion velocity $\\sigma$ \\citep{fm00,g00}. This tight correlation $M_{\\rm bh}-\\sigma$ is widely used to estimate the central black hole masses of active galactic nuclei (AGNs). Unfortunately, the stellar dispersion velocity $\\sigma$ is available only for a small fraction of AGNs. \\citet{md02} derived a very tight correlation between host galaxy absolute magnitude $M_R$ at $R$-band and black hole mass $M_{\\rm bh}$. \\citet{cao03} used this relation $M_{\\rm bh}-M_R$ to estimate the central black hole masses of 29 BL Lac objects. The central black hole masses of three sources in their sample have also been measured from the stellar dispersion velocity \\citep{f02,b03}, which agree well with the black hole masses estimated from the host galaxy absolute magnitude $M_R$. \\citet{o02} also found that the black hole masses estimated from the host galaxy luminosity are quite reliable for radio galaxies. The BLR sizes $R_{\\rm BLR}$ of broad-line ${\\rm H}\\beta$ are measured by \\citet{k00} for a sample of quasars and Seyfert galaxies using reverberation mapping method. They found a tight correlation between the BLR size $R_{\\rm BLR}$ and optical continuum luminosity $\\lambda L_{\\lambda}$. Using the width of the broad emission line and measured BLR size $R_{\\rm BLR}$, they estimated the central black hole masses of the sources in their sample assuming the clouds in BLRs to be virilized. For sources at high redshifts, the emission of line ${\\rm H}\\beta$ is usually unavailable. Instead, the width of the line Mg\\,{\\sc ii} can be used to estimate the central black hole masses \\citep{mj02}. Mg\\,{\\sc ii} is a low-ionization line as ${\\rm H}\\beta$, so that Mg\\,{\\sc ii} is expected to be produced in the same region as ${\\rm H}\\beta$, which is supported by the tight correlation between the FWHM of Mg\\,{\\sc ii} and ${\\rm H}\\beta$ ($V_{\\rm FWHM}$({\\rm Mg\\,{\\rm\\sc ii}})$\\sim V_{\\rm FWHM}({\\rm H}\\beta$)) found by \\citet{mj02}. It is therefore reasonable to expect that Mg\\,{\\sc ii} and ${\\rm H}\\beta$ are produced in the same region. Using the same sample and the BLR sizes measured by \\citet{k00}, \\citet{mj02} obtained a correlation between $R_{\\rm BLR}$ and the monochromatic continuum luminosity at 3000 {\\AA}, which is useful for estimate of black hole masses of the sources at high redshifts with only Mg\\,{\\sc ii} emission line profiles. For most AGNs, their BLR sizes have not been measured directly by reverberation mapping method and the empirical relation $R_{\\rm BLR}-\\lambda L_{\\lambda}$ is used to derive $R_{\\rm BLR}$. Combining the line width, the central black hole masses can be estimated by assuming the motion of clouds in BLRs to be virilized. The estimated black hole masses depend sensitively on the velocity of the clouds in BLRs ($\\propto V_{\\rm BLR}^2$). The broad-line width is mainly governed by the component of the cloud velocity $V_{\\rm BLR}$ projected to the line of sight. If the motion of BLR clouds is anisotropic (e.g., disk-like BLR geometry), the estimate of black hole mass becomes complicated (e.g., Jarvis \\& McLure, 2002). \\citet{md01} argued that BLRs in some AGNs have disk-like geometry. If the disk-like BLR geometry is indeed present, the observed broad-line width depends sensitively on the orientation of disk axis and the orientation o is therefore crucial in the estimate of black hole mass from its broad-line width. In the unification schemes, the jets of BL Lac objects are supposed to be inclined at small angles with respect to the line of sight (see Urry \\& Padovani, 1995 for a review), so the broad-line profiles will be significantly narrowed if disk-like BLRs are present perpendicular to the jets, which can be used to test the geometry of BLRs in BL Lac objects. In this paper, we use the observed broad emission line widths and black hole masses derived from host galaxy luminosity to test the geometry of BLRs in BL Lac objects if the black hole masses can be estimated by independent method. The cosmological parameters $\\Omega_{\\rm M}=0.3$, $\\Omega_{\\Lambda}=0.7$, and $H_0=70~ {\\rm km~s^{-1}~Mpc^{-1}}$ have been adopted in this paper. ", "conclusions": "We plot the relation between the ionizing luminosity $\\lambda L_{\\lambda}(3000)$ at 3000 {\\AA} and BLR size $R_{\\rm BLR}$ in Fig. \\ref{fig1}. The BLR size $R_{\\rm BLR}$ is derived from the width of broad line Mg\\,{\\rm\\sc ii} and black hole mass $M_{\\rm bh}$ on the assumption of isotropic motion of clouds in the BLR, i.e., $f=\\sqrt{3}/2$ is adopted in Eq. (\\ref{rblrmbh}). We can also derive the BLR size $R_{\\rm BLR}^{\\rm emp}$ using the empirical relation (\\ref{rblr3000}). If the motion of BLR clouds is indeed isotropic, one may expect similar BLR sizes derived by these two different methods. However, it is found that the sizes of BLRs $R_{\\rm BLR}$ in all these sources are $\\sim$2-3 orders of magnitude larger than $R_{\\rm BLR}^{\\rm emp}$ expected by relation (\\ref{rblr3000}). We note that only upper limits on the masses of the black holes in 9 of all 16 sources. The BLR sizes may be over-estimated for these 9 sources, because of the BLR sizes $R_{\\rm BLR}$ being derived from the line widths and black hole masses (see Eq. \\ref{rblrmbh}). The black hole masses estimated from the host galaxy luminosity for these 16 sources are in the range of $\\sim 10^{8.6-10.4} M_{\\odot}$ (see Table 1). The deviations of the BLR sizes $R_{\\rm BLR}$ from $R_{\\rm BLR}^{\\rm emp}$ expected by relation (\\ref{rblr3000}) cannot be solely attributed to the overestimate of black hole masses for those nine sources with upper limits on galaxy luminosity, unless the black hole masses have been overestimated by 2$-$3 orders of magnitude, i.e., the realistic black hole masses should be in the range of $\\sim 10^{6-8}$ $M_\\odot$ for these sources, which seems impossible. It will be more difficult to attribute such deviations to the overestimate of black hole masses for those 7 sources with well measured host galaxy luminosity. The black hole mass of the source $1807+698$ has been measured from its stellar dispersion velocity $\\sigma$ \\citep{f02,b03}, which is consistent with our estimate of the black hole mass $10^{8.88} M_{\\odot}$. For this source, its BLR size $R_{\\rm BLR}$ derived from relation (\\ref{rblrmbh}) is about three orders of magnitude higher than $R_{\\rm BLR}^{\\rm emp}$ predicted by relation (\\ref{rblr3000}) between $R_{\\rm BLR}^{\\rm emp}$ and $\\lambda L_{\\lambda}(3000)$. \\figurenum{1} \\centerline{\\includegraphics[angle=0,width=10.0cm]{a0312f1.ps}} \\figcaption{\\footnotesize The relation between the BLR size $R_{\\rm BLR}$ and the ionizing luminosity $\\lambda L_{\\lambda}(3000)$ at 3000 {\\AA} derived from narrow-line luminosity $L_{[{\\rm O}_{\\rm II}]}$. The line represents the correlation $R_{\\rm BLR}^{\\rm emp}-\\lambda L_{\\lambda}(3000)$ defined by Seyfert 1 galaxies and quasars \\citep{mj02}. The source names are labelled in the plot.\\label{fig1}} \\centerline{} The ionizing luminosity at 3000 {\\AA} is derived from the luminosity of narrow-line [O\\,{\\sc ii}] adopting EW$_{\\rm ion}$ =10 {\\AA} \\citep{w99}, which is in general consistent with the optical spectroscopic observations on the radio quasars selected from the Molonglo Quasar Sample (MQS), of which the equivalent widths of [O\\,{\\sc ii}] in the source frame are in the range from less than 1 {\\AA} to more than 100 {\\AA} with an average of 14.7 {\\AA} \\citep{b99}. However, using a single value EW$_{\\rm ion}$ =10 {\\AA} is still a rough estimate, and may induce uncertainties on the estimates of ionizing luminosity. In Fig. \\ref{fig1}, we find that the deviations of $R_{\\rm BLR}$ from $R_{\\rm BLR}^{\\rm emp}$ will not be solved on the isotropic BLR geometry assumption even if a rather small EW$_{\\rm ion}$=1 {\\AA} is adopted, i.e., the ionizing luminosity is an order of magnitude higher than the present values. In Table 1, we have listed the measured equivalent widths of [O\\,{\\sc ii}] in the source frame for the sources in our present sample, which are in 0.2$-$9.2 {\\AA}. Considering the optical continuum emission contributed by the beamed emission from the jets in these BL Lac objects, their equivalent widths (EW$_{\\rm ion}$) corresponding to the ionizing optical continuum emission should be larger than the measured values listed in Table 1. This implies that the uncertainties on the estimates of the ionizing optical continuum luminosity would not change the conclusion on this point. If photo-ionization of the gases in narrow-line region by radiative shocks driven by the radio source is important (e.g., Inskip et al. 2002), the central ionizing luminosity should be lower than the present values, which would lead to even larger deviations. \\citet{wz03} found that the BLR sizes of dwarf AGNs are systematically larger than the prediction of $R_{\\rm BLR}-\\lambda L_{\\lambda}$ correlation defined by Seyfert galaxies and quasars \\citep{k00}. They suggested that the flat ionizing spectra are in these dwarf AGNs as predicted by advection dominated accretion flow (ADAF) models, and the BLRs in those dwarf AGNs have lower ionization or/and lower density than those in Seyfert 1 galaxies and quasars of \\citet{k00}'s sample. The sources in our present sample have brighter ionizing luminosity than those dwarf AGNs. As most BL Lac objects do not exhibit any line emission, all these 16 BL Lac objects have measured broad-line profiles and narrow-line emission have relatively high ionizing luminosity amongst all BL Lac objects \\citep{cao02a}. There is a critical accretion rate $\\dot{m}_{\\rm crit}$, and an ADAF can exist only if accreting at a rate $\\dot{m}<\\dot{m}_{\\rm crit}$(e.g., Mahadevan, 1997), which leads to an upper limit on optical continuum emission from an ADAF for a given black hole mass \\citep{cao02a}. \\citet{cao03} calculated optical spectra of ADAF$+$SD(standard disk) systems and compared them with the observed spectra, which suggests that only ADAFs themselves are unable to produce such bright ionizing optical continuum emission and standard thin disks should be present at least in the outer regions of the disks for these BL Lac objects. The ionizing photos are therefore mainly from the standard thin disk regions in these BL Lac objects, unlike the dwarf AGNs considered by \\citet{wz03}. The luminosity $\\lambda L_{\\lambda}(3000)$ of the sources in the sample used to derive the correlation $R_{\\rm BLR}-\\lambda L_{\\lambda}(3000)$ is in the range of $\\sim 10^{34-39}$ W \\citep{mj02}. The ionizing luminosity of the BL Lac objects in present sample at 3000 {\\AA} are in the range of $\\sim 10^{35.8-38}$ W, which is in the similar range as their sample. The $R_{\\rm BLR}-\\lambda L_{\\lambda}(3000)$ relation derived from a sample of quasars and Seyfert galaxies should be valid for these BL Lac objects, unless the physics of BLRs in these BL Lac objects is significantly different from that in the sources of the sample considered by \\citet{mj02}. We derive the BLR sizes of these BL Lac objects assuming the motion of clouds to be isotropic, which may not be the case in these sources. Our estimate of BLR sizes may strongly be overestimated due to anisotropic cloud motion, if the velocity component projected to the line of sight is only a small fraction of its real velocity. A most likely candidate for such anisotropic motion of clouds is the clouds orbiting in a disk-like BLR (e.g., McLure \\& Dunlop, 2001). For the clouds orbiting in a disk-like BLR, the correction factor $f$ in Eq. (\\ref{rblrmbh}) is $1/(2\\sin i)$, where $i$ is the angle of the axis inclined to the line of sight \\citep{md01}. If this is the case, we can estimate the inclination angle $i$ of these BL Lac objects assuming they indeed to obey the correlation $R_{\\rm BLR}-\\lambda L_{\\lambda}(3000)$ suggested by \\citet{mj02}, i.e., the value of $f$ is estimated by letting $R_{\\rm BLR}=R_{\\rm BLR}^{\\rm emp}$. The derived results are listed in Table 1. We find that the inclination angles are around $\\sim 2^\\circ-12^\\circ$ for these BL Lac objects. There is evidence that the velocity field of BLR is better described by a combination of a random isotropic component, with characteristic velocity $V_{\\rm r}$, and a component only in the plane of the disk, with characteristic velocity $V_{\\rm p}$ \\citep{wb86}. In this case, the observed FWHM will be given by \\be V_{\\rm FWHM}=2(V_{\\rm r}^2+V_{\\rm p}^2\\sin^2 i)^{1/2} \\ee\\citep{md01}, so $f=0.5[(V_{\\rm r}/V_{\\rm p})^2+\\sin^2 i]^{-1/2}$. If the random isotropic component is important, i.e., $V_{\\rm r}$ is comparable with $V_{\\rm p}$, then the term $(V_{\\rm r}/V_{\\rm p})^2$ cannot be neglected and the derived inclined angle of the disk axis will be less than that listed in Table 2. This is in general consistent with the unification schemes that the jets of BL Lac objects are inclined at small angles to the line of sight." }, "0403/astro-ph0403584_arXiv.txt": { "abstract": "{ GRB 020410 is by far the longest $\\gamma$-ray burst (with a duration of about 1600~s) to have been followed up from the X-ray through the radio regime. Afterglow emission was detected in X-rays and at optical wavelengths whereas no emission was detected at 8 GHz brighter than 120 $\\mu$Jy. The decaying X-ray afterglow, back extrapolated to 11 hours after the burst, had a flux of $7.9\\times10^{-12}$ \\fu\\ (2--10 keV); the brightest detected so far. No direct redshift determination is available yet for this GRB, but according to the empirical relationship between the peak energy in the $\\nu F_\\nu$ spectrum and the isotropic energy output, $z$ is constrained in the range 0.9--1.5. The reconstructed optical afterglow light curve implies at least two breaks in the simple power-law decay. This may be related to emergence of a SN, or refreshment of the external shock by a variation in the circumstellar medium. By comparing the backward extrapolation of the 2--10 keV afterglow decay, it is shown that the long duration of the prompt emission is not related to an early onset of afterglow emission, but must be related to prolonged activity of the ``central engine''.} ", "introduction": "Gamma-ray Bursts (GRB) show great diversity with regard to both their durations and spectral properties. GRBs last from a fraction of a second to thousands of seconds, as established by the BATSE survey (e.g. \\cite{Paciesas99}). Prompt X-ray counterparts of GRBs, detected by Ginga, \\sax, HETE-2 have a very wide distribution of intensities and durations. Tails and precursors of X-ray counterparts were also observed by WATCH/Granat (\\cite{ct94}). Those events characterized by an X-rays-to-$\\gamma$-rays (2--10/40--700 kev) fluence ratio larger than $\\sim0.5$ are classified as X--ray rich (e.g. \\cite{Feroci01}). Moreover, transient X--ray sources with characteristics similar to those of GRB counterparts, although with no simultaneous GRB detection (so called ``X-ray flashes''; \\cite{Heise01, moc03}) were detected by the \\sax\\ Wide Field Cameras (WFC) and, subsequently, by the HETE--2/FREGATE instrument. Recently (\\cite{Zand03}), have reported the detection of 4 long, faint X-ray transients during sky surveys with the \\sax-WFC. Three of these are confirmed GRBs, because they coincide with BATSE detections. They show durations ranging from 540 s to 2550 s and are characterized by a mildly soft spectrum. The very different ratios of $\\gamma$-ray vs X-ray peak fluxes or fluences point either to different viewing angles of the relativistic jets in which GRBs are formed (e.g. \\cite{Granot02, Yamazaki02}) or to a different amount of baryon contamination of the fireball (e.g. \\cite{Dermer99, Huang02}). Furthermore, the existence of a class of GRBs with long X-ray durations is important for the investigation of the connection between the prompt and afterglow components and may also suggest a high redshift origin. However, redshift constraints imposed on XRF 020903, 030723 and, possibly, 031203 do not support the high redshift scenario (\\cite{Soderberg03, Prochaska03, Fynbo04}). \\grb, first detected in X-rays only by the \\sax-WFC (\\cite{Gandolfi02}), stands out for its long duration, more than 1500 s in the 2--28 keV band (see Sect. \\ref{wfc_an}), and for the relative weakness of its $\\gamma$-ray signal, detected by Konus-Wind in an offline analysis. Based on its X-ray-to-$\\gamma$-ray fluence ratio (see Sect. 3), \\grb\\ lies in the soft tail of genuine GRBs and marginally qualifies as an ``X-ray rich'' GRB (see \\cite{Heise01}). Upon detection of the GRB we started an X-ray and optical search and monitoring campaign of its afterglow. We present here the results of our study of the prompt and afterglow emission. ", "conclusions": "\\grb\\ is by far the longest GRB event for which X-ray afterglow emission and an optical counterpart (though weak) have been discovered. The long duration of \\grb\\ can be considered ``tail of the distribution'' rather than a peculiar case, though peculiar circumstances are required. As the other very long GRBs detected by the WFC (\\cite{Zand03}), this event shows a high ratio between X-ray and $\\gamma$-ray fluence, although it cannot be classified as an X-ray rich GRB. In addition, apart the high X-ray content, X-ray rich GRBs are also characterized by the lack of optical afterglow (not true for the X-ray flashe 030723 and possibly 020903), which is not the case for \\grb. We note, however, that the two things could not be linked as the rapidly fading OT of the ``not-quite'' X-ray rich GRB 021211 has demonstrated (\\cite{crew03}). Spectroscopically, \\grb\\ does not show peculiarities. This in addition to (a) the measured upper limit on the intrinsic absorption, (b) the ``smooth\" increase of the spectral index, (c) the late time X-ray ``afterglow\" decay that does not connect to the late time prompt emission, suggest the prompt event is all due to internal shock rather than being a superimposition of internal and external shocks, or at least the latter is negligible. In the fireball model, a connection between the long duration of the ``prompt'' event and the slow decay of the afterglow could be accounted for e.g. by the superimposition of several (external) shocks produced by each of the main peak of the internal shock, which are all relatively long when compared to normal GRBs, causing a continuous refreshment of the external shock (e.g. \\cite{bj02}). Unlike what is observed for several GRBs (e.g. \\cite{costa99}) the backward extrapolation of the afterglow fading law of \\grb\\ is inconsistent with the flux measured during the last part of the prompt emission (see Fig.~\\ref{f4_xdecay}). This may be linked to the extremely long duration of the event and prevents us to derive an indication of the afterglow emission onset time. We can estimate an upper limit to the afterglow 2--10 keV fluence if, by following \\cite{Frontera00}, we assume that the afterglow emission starts at 63\\% of the duration of the GRB and thus we integrate the fading law between 973 s and $1\\times10^6$ s. The result is $1.96\\times10^{-6}$ erg cm$^{-2}$, corresponding to about 34\\% of the fluence measured in the prompt event in the same energy range and to about 9\\% of the prompt fluence in 40--700 keV. These values are well within the observed range of normal GRBs (\\cite{Frontera00}). Alternatively, it is possible to identify the onset of the external shock at $t\\sim500$~s when the spectrum of the prompt emission becomes consistent with the late time MECS spectra. In the simple case in which the fireball is homogeneous and thin, the GRB variability should be suppressed and the lightcurve be described as a power-law initially rising as $t^2$ and then smoothly turning over to a decay slope which depends on the spectral range and dynamics of the fireball (\\cite{SaPi99}). In fact the lightcurve of \\grb\\ is highly variable after the spectral transition, showing a prominent emission episode at $t~1500$~s (P4 in Fig.~\\ref{f1_promptlc}). This behavior can be understood if the inner engine does not turn off at the end of the gamma-ray phase, but releases a sizable amount of energy at $t\\sim1500$~s. This late emission, however, should be inefficient in the production of $\\gamma$-rays or, in terms of the internal-external shock scenario, it should avoid the internal shock phase. The time $t\\sim1500$~s is not the deceleration time of the fireball, but the delay with which the inner engine released the fireball component that re-energized the external shock to produce the P4 rebrightening. The cause of the lack of $\\gamma$-ray emission associated with the delayed energy release is not clear and, lacking WFC data for the P4 episode, it is difficult to constrain observationally; though the slight count excess in the Konus soft band could be an hint. The delayed energy release may however be associated to the recycling of the energy wasted while the relativistic jet propagates into the host star (\\cite{Mesz01, Ram02}). In that case, the acceleration of the delayed fireball takes place at the star surface, and is therefore characterized by a variability timescale many orders of magnitude larger than that of the jet, effectively preventing the occurrence of internal shocks. We also note that among the other GRBs afterglows for which the extrapolation of the decay law to the prompt emission is inconsistent with the observed flux, GRB 990704 (the X-ray richest event observed by \\sax) is the only analogous case. The afterglow X-ray flux decay of XRF 031203 also shows an extrapolation below the ``probable'' prompt flux (\\cite{Watson04}). GRB 990510, 010222, 010214 show an extrapolation \\emph{above} the prompt emission, which is explained with a break a few hours after the onset (\\cite{Pian01, Zand01, Guidorzi03}) The peak width dependence as function of the energy was tested for P1 and P3 (see Fig.~\\ref{f1_promptlc}). To this aim we produced rebinned light curves with bin size between 1 and 8 s. Their FWHM were obtained using Gaussian fits; a 20\\% systematic error on their estimate was added. By using a law ${\\rm FWHM}=kE^\\alpha$ (expected by the synchrotron model, Fenimore et al. (1995)) for the two peaks we obtain $\\alpha=-0.48\\pm0.20$ for the P1 and $\\alpha=-0.44\\pm0.12$ for P3 (errors are 90\\% confidence level). These results are consistent with the results from the BATSE GRBs (\\cite{Fenimore95}) as well as for GRB 960720 (\\cite{Piro98}) and 990704 (\\cite{Feroci01}). Due to the paucity of data and the complexity of the optical light curve, it is not possible to constrain the fireball and environment properties completely. From X-ray spectroscopy we infer the electron distribution slope $p=2.1\\pm0.25$ under the assumption that X-rays are above the synchrotron cooling frequency. Due to the large uncertainty, the X-ray decay slope of $\\delta_{\\rm X}=0.81\\pm0.07$ can be accounted for both in an ISM and wind environment. It is tantalizing to note, however, that the early time optical slope seems to be larger than the X-ray one. This would fit in a wind environment scenario, consistent with the possible detection of a supernova bump at late time (\\cite{Levan04}). In this case one would expect $\\delta_{\\rm X}=(3p-2)/4=1.1\\pm0.2$ and $\\delta_{\\rm O}=(3p-1)/4=1.3\\pm0.2$, fully consistent with the X-ray slope and the optical lower limit. Even this interpretation bears however some degree of uncertainty. \\grb\\ has a flux of 10.5 $\\mu$Jy in R and $7.9\\times 10^{-12}$ \\fu\\ in X-rays which falls outside the distribution found by \\cite{depas02}) (see their Fig.~5) and would classify it as a dark GRB (\\cite{laz02}). Even assuming that the cooling frequency lies exactly at the edge of the \\sax\\ band, the synchrotron spectrum would over-predict optical emission by a factor $\\sim5$. There are two possible interpretations for this. One possibility is that the X-ray emission is boosted by an IC component, like in the case of GRB 000926 (\\cite{har01}). This would require a moderately dense environment, either uniform or stratified. Alternatively, the optical emission may be extincted by a sizable amount of dust in the host galaxy, with $A_{\\rm V}\\sim2$. This would correspond, for a Galactic mixture, to a column density $\\nh\\sim3\\times10^{21}$~cm$^{-2}$, consistent with the upper limit derived from X-ray spectroscopy. The lack of constraints on the optical spectrum prevents us to reach a definite conclusion. The optical spectrum should be bluer in the case of IC emission and red in the case of dust obscuration. Assuming that the emission line in the MECS spectra is real and due to fluorescence of H-like iron (rest energy of 6.97 keV), then the change in line position can be explained by a variable iron recombination edge showing its maximum in the second half of ToO~1 (or later, but before ToO~2). In fact if we derive the redshift from the line position in ToO~2, we obtain $z\\simeq1.7$ which leads to a recombination edge of $\\sim3.4$ keV. Also the ratio between the iron recombination edge rest energy, 9.28 keV, and 6.97 keV is $\\simeq 1.3$, like for the ratio of the ToO~1b over ToO~2 line energies. Again, our statistics does not allow us to perform a simultaneous fit for a Gaussian and a recombination edge line. However this hypothesis appears in agreement with the data. As no direct $z$ measurement exists for \\grb, we calculated the peak energy $E_{\\rm p}$ in the $\\nu F_\\nu$ spectrum and the isotropic energy $E_{\\rm rad}$ for a grid of $z$ values. We then compared the results with the relation reported by Amati et al. (2002). We find that the relation is satisfied (with a discrepancy level $<20\\%$) for $0.9 < z < 1.5$ and $1.1\\times10^{53}< E_{\\rm rad} < 3.0\\times10^{53}$ erg. This range of $z$ would exclude the value of 0.5 obtainable by assuming a 1998bw-like SN re-bump (\\cite{Levan04}) and is marginally in agreement with the value derived above. Even assuming that our flux estimate for the missing part of the X-ray light curve must be increased by an extra 20\\% (which is unlikely), the lower limit for $z$ becomes 0.6. Besides this, we note that $z\\simeq0.5$ together with the reported magnitude of ${\\rm V}\\simeq 28.7$ for the host galaxy (\\cite{Levan04}) would place it in the very low end of the galaxy luminosity function (${\\rm M_{\\rm V}} = -14.3$), which is unusual for GRB hosts; this independently of considering the X-ray spectrum derived $\\nh\\simlt3\\times10^{21}$~cm$^{-2}$ as being ``local'' or ``global''." }, "0403/astro-ph0403067_arXiv.txt": { "abstract": "We summarize some recent results from our observational campaign to study the central regions of spiral galaxies of late Hubble type (Scd - Sm). These disk-dominated, bulgeless galaxies have an apparently uneventful merger history. The evolution of their nuclei thus holds important constraints on the mechanisms that are responsible for bulge formation and nuclear activity in spiral galaxies. We discuss the structural properties, stellar populations, and dynamical masses of the compact, luminous star cluster that is found in the nuclei of most late-type spiral galaxies. Although preliminary, our results strongly indicate that many galaxies of our sample experience repeated periods of nuclear star formation. While the exact mechanism that leads to the required high gas densities in the galaxy nucleus remains unclear, results from our recent CO survey of late-type spirals demonstrate that in most cases, the central kpc contains enough molecular gas to support repetitive nuclear starbursts. ", "introduction": "In most formation scenarios for spiral galaxies, the central bulge is the ``trashbin of violent relaxation'' where a dynamically hot stellar component has formed either through external potential perturbations such as early mergers of proto-galaxies (e.g. Carlberg 1992), or perhaps via internal effects such as violent bar instabilities (e.g. Norman, Sellwood, \\& Hasan 1996). The latest-type spirals, then, must have lived very sheltered and uneventful lives, since their central ``trashbin'' is virtually empty, i.e. they are devoid of prominent starburst events, have no discernible stellar bulges, and rarely show signs of nuclear activity. These galaxies often have gently rising rotation curves (e.g. Matthews \\& Gallagher 1997) that indicate a nearly homogeneous mass distribution on scales $\\sim 1\\kpc$. On these scales, gravity therefore hardly provides a vector pointing at the center, and it is not obvious that the nucleus of these galaxies is well-defined and a unique environment. \\begin{figure}[!ht] \\vspace*{4cm} \\centerline{see attached file fig1a.gif} \\plotone{fig1b.ps} \\caption{Top: HST/WFPC2 F814W (I-band) images of three representative nuclear star clusters. Shown is the PC chip with a field of view of $\\approx 35\\arcsec \\times 35\\arcsec$. The bar in the top left of each panel denotes a spatial scale of 1 kpc, the north-east orientation is indicated by the compass arrow. Bottom: I-band surface brightness profiles, measured from elliptical isophote fits to the images above. Note the clear transition between the underlying disk and the NC at radii around $0.2\\arcsec$. Also shown are analytical fits to the disk and cluster profiles that yield the cluster photometry (for details, see B\\\"oker et al. 2002.)} \\label{fig:images} \\end{figure} Surprisingly, the photocenter of many late-type spiral galaxies is nevertheless occupied by a compact, luminous stellar cluster (Phillips et al. 1996; Carollo et al. 1998; Matthews et al. 1999; B\\\"oker et al. 2002, hereafter Paper~I). Figure~\\ref{fig:images} shows HST/WFPC2 I-band images of three representative examples of such nuclear star clusters (NCs), together with their surface brightness profiles as measured from elliptical isophote fits to the WFPC2 data. The typical luminosities of NCs are in the range $10^6$ - $10^7\\,\\lsun$ (Paper~I). Nuclear clusters are therefore much brighter than average stellar clusters in the disks of nearby spiral galaxies (e.g. Larsen 2002), and comparable to young ``super star clusters'' in luminous merger pairs (Whitmore et al. 1999) or circumnuclear starforming rings in spiral galaxies (e.g. Maoz et al. 2001). Knowledge of the stellar populations and masses of NCs is essential in order to constrain the mechanism(s) that lead to their formation. The age(s) of the stellar population(s) can be determined from spectral synthesis methods applied to medium-resolution spectra of NCs. Measuring the stellar masses of NCs requires both accurate knowledge of the velocity dispersion (from high-resolution spectra) and the cluster light distribution (from HST imaging). This kind of analysis has been successfully applied to NCs (B\\\"oker et al. 1999) as well as young clusters in the disks of nearby galaxies (Smith \\& Gallagher 2001; Mengel et al. 2002, McCrady et al. 2003). In this paper, we summarize our ongoing study of the properties and formation mechanism(s) of NCs in late-type spirals. ", "conclusions": "In ellipticals and ``classical'' spirals, phenomena that indicate nuclear activity such as AGN and massive black holes have received much attention over the past decades. Only recently, however, has it become evident that even in bulgeless, ``pure'' disk galaxies which are generally devoid of any obvious signs for nuclear activity, the galaxy center is a ``special'' location in that it is occupied by a massive, compact, and often young stellar cluster. These nuclear star clusters (NCs) have sizes and luminosities that are very similar to those of young super star clusters (SSCs) found in the disks of starburst galaxies. However, NCs most likely have a different, more complex, formation history, as can be inferred from population synthesis age-dating of their spectra as well as estimates of their masses from stellar dynamical modelling. Unless one is willing to believe that we live in a special time, the fact that many nuclear star clusters are young suggests that they experience repetitive ``rejuvenation'', most likely due to infall of molecular gas into the central few pc and associated star formation. This is not implausible, because bulgeless spirals are in many ways normal spirals. In particular, their molecular gas content follows the same scaling relation with galaxy luminosity as earlier-type spirals, and typical molecular gas masses in their central kpc are of the order $10^7\\,\\msun$. While the ``duty cycle'' for nuclear starbursts still needs to be reliably established - which is one of the goals of our project - the possibility of repetitive nuclear cluster formation in late-type spirals has interesting consequences for galaxy evolution scenarios, both with respect to morphological classification (i.e. Hubble type) and nuclear activity (i.e. black hole growth)." }, "0403/astro-ph0403317_arXiv.txt": { "abstract": "We introduce a model for pulsars in which non-radial oscillations of high spherical degree (\\el) aligned to the magnetic axis of a spinning neutron star reproduce the morphological features of pulsar beams. In our model, rotation of the pulsar carries a pattern of pulsation nodes underneath our sightline, reproducing the longitude stationary structure seen in average pulse profiles, while the associated time-like oscillations reproduce ``drifting subpulses''---features that change their longitude between successive pulsar spins. We will show that the presence of nodal lines can account for observed $180^\\circ$ phase jumps in drifting subpulses and their otherwise poor phase stability, even if the time-like oscillations are strictly periodic. Our model can also account for the ``mode changes'' and ``nulls'' observed in some pulsars as quasiperiodic changes between pulsation modes of different (\\el) or radial overtone ($n$), analogous to pulsation mode changes observed in oscillating white dwarf stars. We will discuss other definitive and testable requirements of our model and show that they are qualitatively supported by existing data. While reserving judgment until the completion of quantitative tests, we are inspired enough by the existing observational support for our model to speculate about the excitation mechanism of the non-radial pulsations, the physics we can learn from them, and their relationship to the period evolution of pulsars. ", "introduction": "\\label{intro} Upon the discovery of radio pulsations from pulsars by \\citet{hew68}, \\citet{rud68}~ proposed that the pulses arose from non-radial oscillations of a neutron star. This idea was quickly displaced by a rotational model \\citep{gol69}, but \\citet{dra68}~ again raised the possibility of pulsations when they measured individual pulse sequences for two pulsars and found within them narrow subpulses that moved to successively earlier times within the main pulse. Because this drift represented the presence of a ``second periodicity'' incommensurate with the spin period, it was natural to propose a time-like oscillation of the star. Subsequent measurements, however, revealed complex subpulse patterns that did not conform to a pulsation model in any obvious way. Moreover, the persistence of unique subpulse shapes from pulse to pulse, along with problems of phase stability we will address in later sections, led Drake to conclude that the drifting subpulses were incompatible with the pulsation hypothesis \\citep[see][]{sta70,hew70}. Ultimately, pulsations were abandoned in favor of purely geometric models, although they reappeared from time-to-time in the theoretical literature \\citep[notably][]{hans80,vanh80,mcd88first,car86,fin90,rei92,str93}. Most recently \\citet{dun98}~ invoked toroidal modes to account for oscillations of soft gamma repeaters, but other than the work of \\citet{str92}~ and \\citet{scvfirst92}, there has been no determined attempt to account for the properties of classical pulsars with models involving non-radial pulsations. Instead, most current models, though not all \\citep[cf.][]{lyn88,han01}, incorporate a circulating pattern of sub-beams, whose motion about the magnetic pole produces the drifting subpulses. In these models, pulsar emission comes from accelerated particles that originate near the pulsar magnetic pole and travel along curved paths in the star's magnetic field \\citep[see][]{rad69,kom70}. The radiation is confined to a narrow beam by the dipole magnetic field geometry \\citep{gj69} and relativistic beaming along the direction of particle motion, which is roughly parallel to the magnetic axis, not perpendicular as in the models of \\citet{gol69}, \\citet{smi70} and \\citet{zhe71}. The observed brightness of pulsar beams effectively demands that the radiation is coherent, but the question of how it is produced is not settled \\citep{jes01,les98,mel95}. Early studies of pulsar single pulses and average pulse shapes \\citep[][and others]{tay75first, lyn71}~ led to the addition of more elaborate emission structures within the model pulsar beam. These features sweep past our sightline and recreate the variety of pulse shapes we observe. \\citet{bac76}~ described a target-shaped emission pattern (a central core surrounded by an annulus) that can reproduce a wide variety of pulse morphologies depending on whether our sightline crosses the center of the pattern, yielding a three component pulse, or crosses only the annulus, resulting in a one or two component pulse. \\citet{ost77} added a second annulus and rotating features to reproduce pulses with more than three components and drifting subpulses. In 1975, \\citet{rud75} supplied a physical basis for the model by suggesting that the emission arises from localized discharges or sparks near the polar cap. These are arranged in annular patterns, and rotate naturally due to the crossed components of the magnetic and electric fields. In addition to the fixed and drifting substructure, models must account for observations of two kinds of discrete events observed in some pulsars; ``mode changes'', which abruptly alter the character of the substructure, and pulse ``nulling'', during which the pulse emission drops below detectable levels for one or more spin periods of the pulsar \\citep{bac70a,bac70b,bar82}. In the \\citet{rud75} model, mode changes and nulling result from a collapse or reorganization of the fixed and moving spark structures, after which they must reappear with the same features they had previously. Several reviewers have summarized observational and theoretical progress in the study of pulsar beams. The most ambitious is Rankin \\citep{ran83a,ran83b,ran86,ran90,rad90,ran93,mit02}, who has both reviewed and synthesized the observations into an empirical model incorporating polarization and spectral behavior. \\citet{man95} gives a somewhat different view of the beam geometry. Most recently, \\citet{gra03} has published a succinct review that includes both ``normal'' and millisecond pulsars. Against this backdrop, as a student project, we conducted a re-analysis of archival data on PSR0943+10 to look for evidence of non-radial pulsations, which, according to theory, might have periods ranging from milliseconds to seconds \\citep{mcd88}. Our analysis, which will appear in a subsequent paper, convinced us that time-like oscillations with a period of $31.8$ msec are a viable alternative to the rotating carousel of emission beams proposed by \\citet{des01}, but we could find no compelling reason other than aesthetics to~{\\it prefer} a pulsational model. In search of a definitive test, we reviewed the extensive observational literature on pulsars, and found intriguing evidence for non-radial pulsations as a universal mechanism for drifting and stationary subpulses. Moreover, we found that the original reason for abandoning pulsational models does not apply to non-radial pulsations of high azimuthal degree (\\el) in which our sightline crosses pulsation nodal lines. The presence of nodal lines increases the variety and subtlety of expected subpulse behavior. The purpose of this paper is to introduce a model in which high \\el~ pulsations aligned to the pulsar magnetic pole take the place of the fixed and moving structures of the circulating spark model, but other details of the geometry remain unchanged. In this paper we will explore only the phenomenological consequences of this substitution, and compare them qualitatively to published observations. We will not discuss in any detail problems in the physics of pulsed radio emission or polarization mechanisms. In \\S\\ref{morph}, we will present the basic features of our model, and explore its observational properties, some of which are not immediately obvious. Our main purpose is to lay the groundwork for future application of the model to radio measurements of individual and average pulse profiles. In \\S\\ref{comp} we will examine qualitative evidence in favor of our model, reserving quantitative comparisons for subsequent papers. The strongest evidence we will present comes from published measurements that show subpulse phase behavior difficult or impossible to explain using the circulating spark model, but demanded by high \\el~ pulsations. We will also discuss analogies between pulsar behavior and that of known pulsating stars, specifically the rapidly oscillating peculiar A stars (roAp) and the pulsating white dwarf stars. This will demonstrate that there are precedents for the model behavior we propose. In \\S\\ref{disc}, we will speculate about theoretical aspects of our model, such as the pulsation driving mechanism, and we will introduce the notion of ``horizontal mode trapping'', which can account for the high \\el~ character of the proposed modes and relate them to the observed period evolution of pulsar beam widths. We will end by highlighting the potential for neutron star seismology, which can yield direct measurements of interesting physical quantities like the buoyancy of neutron star surface oceans. ", "conclusions": "\\label{conc} Whether or not the foregoing discussion has revealed anything about pulsars, it has certainly demonstrated \\citet{cle83} maxim that we can get ``wholesale returns of conjecture out of such a trifling investment of fact.'' Nevertheless, the fact remains: pulsar beams show subpulse phase reversals at the longitude-stationary boundaries separating individual pulse components. We have shown that these changes are comprehensible in the context of an oblique pulsator model incorporating non-radial pulsations of high degree \\el. The important features of our model are: non-radial oscillations aligned to and symmetric about the pulsar {\\it magnetic} axis; surface displacements that follow a spherical harmonic distribution; radio emission that follows the displacements but is never negative; pulsation modes of sufficiently high \\el~ that nodal lines often appear in the pulse window; and pulsation frequencies that remain coherent over many pulsar spin periods. Variations on this basic model might include multiple pulsation modes with non-zero azimuthal orders, pulsations that are distorted, in reality or in appearance, by non-dipole fields, and modes that interact either through mode coupling or a non-linear emission mechanism. Our model qualitatively reproduces the mean shapes of pulsar beams and the radio frequency dependent behavior of subpulses with a minimum of free parameters. In the most basic form of the model these are $\\alpha$, $\\beta$, \\Ptime, \\el, and \\Pone. Our model also dictates specific requirements that can be tested quantitatively using new or archival data. We have embarked on a program to conduct such tests and we encourage others to do likewise. If the model survives these tests, then we will have the opportunity to measure fundamental properties of matter in a domain not accessible to laboratory experiments. The first challenge will be to determine the site of the pulsations, and then to connect measured eigenfrequencies with the eigenmodes of a structural model. Given the number of modes in the pulsation spectrum at large \\el, this may be a daunting task, but even rough identification will provide limits on the thermal, electrical, and mechanical properties of constituents of a neutron star, the densest objects accessible to direct observational scrutiny." }, "0403/astro-ph0403121_arXiv.txt": { "abstract": "We analyze high resolution spectra of a multi--cloud weak [defined as $W_r({\\MgII}) < 0.3$~{\\AA}] absorbing system along the line of sight to PG~$1634+706$. This system gives rise to a partial Lyman limit break and absorption in {\\MgII}, {\\SiII}, {\\CII}, {\\SiIII}, {\\SiIV}, {\\CIV}, and {\\OVI}. The lower ionization transitions arise in two kinematic subsystems with a separation of $\\simeq 150$~{\\kms}. Each subsystem is resolved into several narrow components, having Doppler widths of $3$--$10$~{\\kms}. For both subsystems, the {\\OVI} absorption arises in a separate higher ionization phase, in regions dominated by bulk motions in the range of $30$--$40$~{\\kms}. The two {\\OVI} absorption profiles are kinematically offset by $\\simeq 50$~{\\kms} with respect to each of the two lower ionization subsystem. In the stronger subsystem, the {\\SiIII} absorption is strong with a distinctive, smooth profile shape and may partially arise in shock heated gas. Moreover, the kinematic substructure of {\\SiIV} traces that of the lower ionization {\\MgII}, but may be offset by $\\simeq 3$~{\\kms}. Based upon photoionization models, constrained by the partial Lyman limit break, we infer a low metallicity of $\\sim 0.03$ solar for the low ionization gas in both subsystems. The broader {\\OVI} phases have a somewhat higher metallicity, and they are consistent with photoionization; the profiles are not broad enough to imply production of {\\OVI} through collisional ionization. Various models, including outer disks, dwarf galaxies, and superwinds, are discussed to account for the phase structure, metallicity, and kinematics of this absorption system. We favor an interpretation in which the two subsystems are produced by condensed clouds far out in the opposite extremes of a multi--layer dwarf galaxy superwind. ", "introduction": "\\label{sec:intro} Quasar absorption line systems, as traced by their {\\MgII}, provide a unique way to study our universe. Current classifications schemes for systems with detected {\\MgII} file them into three main categories: those that are characterized by their strong {\\MgII} absorption, those that reveal weak, narrow, single--cloud {\\MgII} profiles, and those that exhibit weak, multiple cloud {\\MgII} absorption. In this paper we analyze in detail the absorption profiles from a particular multiple cloud weak {\\MgII} absorber at $z=1.04$ along the line of sight toward PG~$1634+706$. The longer--term goal of a collection of similar studies will be to understand the relationships between the different classes of absorption systems, and the connections with the different types of galaxies and structures at various redshifts. Although there are no distinct divisions between the three categories of {\\MgII} absorption systems, by their properties they appear to be related to three different types of gaseous structures at $z\\sim1$. Strong {\\MgII} absorbers [those with $W_r(2796)>0.3$~{\\AA}] show Lyman limit breaks and contain multiple clouds spread over tens to hundreds of kilometers per second \\citep{archiveI}. A large majority of these absorbers are known to be associated with luminous galaxies ($>0.05 L^*$, where $L^*$ is the Schechter luminosity), within an impact parameter of $40h^{-1}$~kpc of the quasar \\citep{bb91,bergeron92,lebrun93,sdp94,steidel95,3c336}. All of the strong {\\MgII} absorbers ($dN/dz = 0.91 \\pm 0.1$ at $=0.9$ \\citep{ss92}) can be accounted for by regions of this size around the known population of luminous galaxies. The strong {\\MgII} absorber spectral profiles, in chemical transitions of low and high ionization states, generally require multiple phases of gas, i.e. regions of differing densities that are spatially distinct (e.g. \\citep{ding1634,ding1206}). The kinematics of the low ionization gas is generally consistent with what one would expect from the combined disks and halos of galaxies of a variety of morphological types \\citep{kinmod,steidel02}. In contrast to the strong {\\MgII} absorbers, the single--component systems with weaker {\\MgII} (rest frame equivalent widths $W(2796) <0.3$~{\\AA}) are typically not known to be directly associated with luminous galaxies (i.e. they are not within $50$--$100$~kpc of $>0.05 L^*$ galaxies) \\citep{weak1}. These single--component weak {\\MgII} systems have a significant absorption cross-section, with $dN/dz = 1.10 \\pm0.06$ for $0.02< W< 0.3$~{\\AA} at $0.41$), could also have an origin in superwinds \\citep{bond}. These absorbers have a characteristic saturated ``double--trough'' absorption profile in {\\MgII}, which breaks up into multiple clouds in other weaker low-ionization transitions. As with our two subsystems, the saturated troughs are typically separated by $\\sim100$--$200$~{\\kms}. Strong low-ionization absorption could arise from clouds close to the point of origin of the wind or from early evolutionary stages of the superwind phenomenon. As we reach larger distances or later stages when low--ionization superwind clouds may have fragmented or dispersed, we might expect weaker absorption profiles like the ones we are seeing in the $z=1.04$ absorber toward PG~$1634+706$. If more multiple cloud, weak {\\MgII} absorbers with two subsystems are observed it should be possible to test between an origin in two structures and an origin in a wind or outflow. If winds are responsible then we would expect to see broad, high ionization absorption to the blue of the blueward subsystem and to the red of the redward subsystem. If two separate structures were responsible, the offset low and high ionization components would be due to kinematic differences between two components in such structures. In such a case, the broad, high ionization absorption should arise at random to the red or to the blue of each of the subsystems, and sometimes would even be superimposed. {\\it There are about two thirds as many multiple cloud weak {\\MgII} absorbers as there are strong {\\MgII} absorbers---} There must be a set of structures responsible for multiple cloud weak {\\MgII} absorption with a cross section comparable to the set of $\\sim30$~kpc regions around all $L^*$ galaxies. If these are outer disks around all $>0.05L^*$ galaxies, an annulus with thickness of $\\sim 10$~kpc would be required. \\citet{bond} found that it was plausible for the expected number of superwinds at redshift $z\\sim1.5$ to account for the observed number of saturated ``double -- trough'' strong {\\MgII} profiles. Similarly, we speculate whether the outer winds of a plausible number of bursting dwarfs could account for the observed number of multiple-cloud weak {\\MgII} absorbers. There are about 2/3 as many multiple cloud weak {\\MgII} absorbers are there are strong {\\MgII} absorbers \\citep{weak1,weak2}. Strong {\\MgII} absorption arises from regions around $>0.05 L^*$ galaxies with sizes of $\\sim30h^{-1}$~kpc \\citep{bb91,bergeron92,lebrun93,sdp94,steidel95,3c336}. If the regions around starbursting dwarfs that give rise to multi-cloud weak {\\MgII} absorption have sizes of $10h^{-1}$~kpc, then a number density $6$ times that of $\\sim L^*$ galaxies could account for all these multi-cloud weak {\\MgII} absorbers. Realistically, it seems likely that there would be contributions to multiple cloud weak {\\MgII} absorption from a variety of types of structures and processes. Conversely, all types of structures that contain gas must all make some contribution to the absorber population, contributing either strong or weak {\\MgII} absorption and/or contributing to {\\CIV} or {\\OVI} absorption. It is fairly well understood that $L^*$ galaxies produce most all of the observed strong {\\MgII} absorption at $z\\sim1$. But it is not well understood what type of absorption systems quiescent dwarf galaxies or dwarf galaxies with winds would produce. Above, we have considered the various properties of the $z=1.04$ absorber and discussed what types of structures and processes might be consistent. Clearly, there is no unique interpretation. However, a model in which this absorber is related to dwarf galaxy winds is consistent with most of the listed properties. It can explain the phase structure of the absorber, the relative metallicities of the phases, and the kinematics of the two subsystems relative to the broad, higher ionization components. This model is also appealing in that it accounts for some of the absorption cross section presented by starbursting dwarfs, which were common at redshift one. Many more multiple cloud, weak {\\MgII} absorbers must be studied in order to evaluate the relative contributions of dwarf galaxy winds, of quiescent dwarf galaxies, of outer galaxy disks, and of other phenomena. In conjunction with such a statistical study, it will be essential to find analogs in the local universe so that the host galaxies and responsible processes can be directly identified." }, "0403/hep-th0403270_arXiv.txt": { "abstract": "We consider warped compactifications in $(4+d)$-dimensional theories, with four dimensional de Sitter $dS_4$ vacua (with Hubble parameter H) and with a compact internal space. After introducing a gauge-invariant formalism for the generic metric perturbations of these backgrounds, we focus on modes which are scalar with respect to $dS_4$. The physical eigenmasses of these modes acquire a large universal tachyonic contribution $-12d/(d+2) H^2$, independently of the stabilization mechanism for the compact space, in addition to the usual KK masses, which instead encode the effects of the stabilization. General arguments, as well as specific examples, lead us to conjecture that, for sufficiently large dS curvature, the compactified geometry becomes gravitationally unstable due to the tachyonic growth of the scalar perturbations. This mean that for any stabilization mechanism the curvature of the dS geometry cannot exceed some critical value. We relate this effect to the anisotropy of the bulk geometry and suggest the end points of the instability. Of relevance for inflationary cosmology, the perturbations of the bulk metric inevitably induce a new modulus field, which describes the conformal fluctuations of the 4 dimensional metric. If this mode is light during inflation, the induced conformal fluctuations will be amplified with a scale free spectrum and with an amplitude which is disentangled from the standard result of slow-roll inflation. The conformal 4d metric fluctuations give rise to a very generic realization of the mechanism of modulated cosmological fluctuations, related to spatial variation of couplings during (p)reheating after inflation. ", "introduction": "De Sitter or quasi-de Sitter 4d geometries describe the present day acceleration of the Universe as well as the inflationary expansion at very early times. This stimulates significant interest towards the construction of 4 dimensional dS geometry in the context of fundamental string/M theory, which are formulated in 10/11 dimension, with a compact internal space \\cite{Mal,ds1,KKLT}. Similarly, compactification to 4d dS geometry takes place in phenomenological braneworld models, where the inner space is periodic with orbifold branes at the edges \\cite{ds2}. Although most of the activity in this area has been enhanced by very recent progress (both on the observational and theoretical side) the issue of dimensional reduction to the outer cosmological space--time was popular since the 1980s, either in high dimensional supergravity theories or on phenomenological grounds. For example, see the collection of references on KK cosmology given in \\cite{ss}. The geometries we discuss here have $d$ spatial dimensions wrapped on a compact manifold ${\\cal M}\\,$, in addition to the standard $(3+1)$ space--time. Many of the mentioned examples are covered by the $(4+d)$ dimensional geometry with the metric \\begin{eqnarray}\\label{blocks} d s^2 = e^{2\\hat A(y)}\\left(-dt^2+e^{2Ht} d \\vec x^2\\right)+e^{-2\\hat A(y)}{\\hat g}_{a b} \\left( y \\right) d y^ad y^b \\,, \\end{eqnarray} where the outer space is a dS 4d metric with a Hubble parameter $H$, while ${\\hat g}_{a b}$ is the metric of the compact inner space with $d$ coordinates $y^a$. For generality, we also include the warp factor $\\hat A(y)$. In the following we use greek letters to describe the $(3+1)$ outer space--time coordinates ($\\mu=0,...,3$) while roman letters span the inner compact space coordinates only. Capitalized roman letters span all coordinates. With the re-definition of the inner space metric ${ g}_{a b} \\equiv e^{-4\\hat A(y)}{\\hat g}_{a b}$ and the warp factor $A(y) \\equiv e^{\\hat A(y)}$, we can re-write (\\ref{blocks}) in a conformally-factorized form \\begin{equation}\\label{conf} d s^2=A(y)^2 \\left(d s_4^2 +{ g}_{a b} d y^a \\, d y^b \\right) \\ , \\end{equation} where $ds_4^2$ is de Sitter metric. This form of the metric is more convenient for developing the formalism of metric perturbations around a warped $dS_4 \\times {\\cal M}\\,$ background, which we will present below. >From the string theory/supergravity perspective, recent studies have concentrated on the compatibility of high $(4+d)$ dimensional geometries ($d=6$) with a 4d de Sitter geometry as the outer space--time, and on the stabilization of the internal space to $ dS_4$ \\cite{KKLT}. The bulk geometry, which is usually treated in the supergravity limit, requires a careful study of the $(4+d)$ dimensional (bulk) Einstein equations. The progress in finding solutions is related to the identification of various possible sources for the bulk stress energy tensor $T^{A}_{B}$, including supergravity lagrangian fields, branes with fluxes, etc. In phenomenological braneworld models, the $(4+1)$ or $(4+2)$ dimensional bulk geometry is stabilized, for instance, by means of bulk scalar fields as in the Goldberger-Wise model~\\cite{GW}, or more generally through the Casimir effect \\cite{Cas,pp}. Very often, however, the stabilization is studied at the level of a 4 dimensional effective theory, where the inner space geometry emerges in terms of moduli fields. If the effective potential of the moduli has a minimum, the moduli are considered to be stabilized. An example will be the stabilization mechanism with quantum field theory effects which emerge from the properties of ${\\cal M}\\,$, e.g. the Casimir effect from compact inner dimensions. It is a reasonable assumption that in principle there shall be a bulk prototype $\\langle T^{A}_{B}\\rangle$ (may be as a complicated functional of the metric and the topology) which generates corresponding terms of the 4d effective potential. The version of string theory dimensional reduction of~\\cite{KKLT} is realized with the participation of the instanton effects (which are manifested in 4dim superpotential of $N=1$ supergravity). An interesting question is whether it is possible to think about a bulk prototype $T^{A}_{B}$ of these effects. While it is reasonable to assume that the low energy 4d effective action is Einstein gravity plus moduli fields (or Brans-Dike gravity plus moduli), this picture lacks the full higher dimensional solution of the Einstein equations with proper sources. In particular, the four dimensional description is inadequate for studying the high energy regime (high dS curvature) where the stabilization can break down. For example, in the 5 dimensional braneworld models, the issue of stabilization due to the bulk scalar field can be formulated fully in terms of the stability of the 5 dimensional warped geometry against scalar metric perturbations. From this study one can determine the modes which enter in the exact low energy 4 dimensional description, namely the radion and the other KK modes of the system \\cite{radion}, which cannot be found from an heuristic 4 dimensional effective potential. We are not concerned here with the details of the stabilization mechanism. Our main goal is to determine the effects of the dS expansion on the stability of these geometries. As we will show, compactifications to a 4dim dS space--time are more difficult to achieve than compactification to a 4dim flat space--time. This is due to a tachyonic contribution to the square of the mass of scalar metric perturbations arising from the de Sitter curvature. In what follows we will discuss the properties of $(4+d)$ dimensional classical Einstein equations assuming a bulk $T^{A}_{B}$ as a source, but without specifying it. The gravitational instability which we will discuss comes from the gravity sector, so that the exact form of $T^{A}_{B}$ will not be crucial. Of course it will be attractive to check how the effect works for each particular model of $T^{A}_{B}$. However, we will try to argue that the instability effect is generic. Previous studies of the gravitational sector ({\\it i.e.} metric perturbations) of the compactification concentrated on the dimensional reduction to 4dim flat space--time \\cite{ss}. The instability effect which we find emerges when the outer space is curved instead of flat, and it is proportional to the curvature. Notice that test scalar fields propagating in $dS_4 \\times {\\cal M}$ do not exhibit tachyonic masses under KK projection~\\cite{KS}. Thus, there is a significant difference in the mass spectrum between the test fields and the self-consistent treatment of the metric fluctuations around this background. The plan of this paper is as follows. In Section ~\\ref{sec:metric} we introduce a generic formalism of metric fluctuations around the $dS_4 \\times {\\cal M}$ background and define scalar, vector and tensor fluctuations acording to their transformation properties relative to $dS_4$. In Section~\\ref{sec:lin} we introduce the linearized equations that govern the evolution of scalar perturbations of the metric and show how the tachyonic instability of the scalar modes arises from the gravitational sector. In Sections~\\ref{sec:ex1} and ~\\ref{sec:ex2} we show two explicit example where the instability is manifested. In Section~\\ref{sec:mod} we discuss how the instability may be relevant for the generation of the primordial perturbations, providing a very natural and model-independent realization of the idea of modulated perturbations~\\cite{modulated,dgz}. Conclusions and a brief discussion can be found in Section~\\ref{sec:end}. ", "conclusions": "~\\label{sec:end} A typical denominator in many extensions of the Standard Model is the presece of extra-dimensions. In the simplest possibilities, the extra space is a compact and static manifold ${\\cal M}\\,$. A stabilization mechanism is typically required to explain why ${\\cal M}$ should remain static, while the $3$ noncompact spatial dimensions are undergoing cosmological expansion. In particular, significant activity has been recently made trying to reconcile this picture with a de Sitter (or quasi de Sitter) geometry for the noncompact coordinates. Indeed, observations are telling us that the expansion of the universe was accelerating at very early times, and it is accelerating also at present. The present note is focused on the effects of the inflationary expanison on the stability of ${\\cal M}\\,$. We have argued that the dS curvature (in other words, a nonvanishing expansion rate $H$) has typically the effect to destabilize the internal space, and that any given stabilization mechanism can be effective only up to a certain curvature. We have aimed to discuss this effect in the most general way possible, enlighting the consequences of making the only assumption that the geometry is of the $dS \\times {\\cal M}_4$ type, with an arbitrary compact space of $d$ dimsensions, and possibly with the presence of a warp factor. Such a set-up is also relevant for string-theory, in the supergravity limit. To discuss the stability of the system, we have studied the most general set of perturbations of the $dS_4 \\times {\\cal M}_4$ geometry. It is very convenient to classify the perturbations into irreducible representations of the $dS_4$ symmetry group. The big advantage of doing so, is that modes belonging to different representations are not coupled at the linear level. The perturbations can be devided into scalar, vecor, or tensor modes with respect to the $dS_4$ isometries. After a general classification, we have focused on the scalar modes, since they are ususally the most relevant one for inflationary geometries, and since they encode the effects of the instability we want to discuss. The linearized Einstein equations for the perturbations show that the physical mass squares $m^2$ of these modes acquire a negative contribution due to the $dS$ expansion, \\begin{equation} m_n^2 \\left( H \\right) = - \\frac{12 \\, H^2}{1+d/2} \\,\\,. \\label{tacconc} \\end{equation} This is true for arbitrary ${\\cal M}\\,$, $d\\,$, and for any possible underlying stabilization mechanism. The presence of this tern is a signal for a possible gravitational instability of the system. Indeed, if the whole $m_n^2$ turns out to be tachyonic, the $dS \\times {\\cal M}$ is unstable. Clearly, verifying whether (and at which $H$) the instability takes place is very model dependent issue which should be verified case by case (namely, for any given source $T^A_B \\,$). Indeed, other contributions to $m^2$ also dpend on $H$ (although in most cases only indirectly, due to the fact that these terms depend on other background quantities, and that these quantities are related to $H$ through the background Einstein equations). However, the contribution~(\\ref{tacconc}) is {\\it generally} present, and it has to be taken always into account. Moreover, the study of several different examples, as well as general arguments, lead us to conjecture that the term~(\\ref{tacconc}) is the dominant one at large $H\\,$, and that no stabilization mechanism can be effective up to arbitrarly large dS curvature. In this work, we have presented explicit and exact calculations in two specific examples. The first of them is codimension one braneworld configurations with bulk scalar field(s). As shown in~\\cite{FK}, in this case it is possible to derive an explicit upper bound $m^2 \\leq - 4 \\, H^2 + \\mu_0 \\left( A \\right)\\,$ (where $A$ is the warp factor, see eq.~(\\ref{mass3})) for the mass of the lightest eigenmode. This bound allows to determine at which $H$ the system becomes unstable. The second example is when ${\\cal M}$ is a $d-$sphere, and when the only source term isw a cosmological constant. This system is well known to be unstable, and we have shown that the instability is precisely due to the tachyonic nature of the scalar modes which we have identified in the general calculation. There are also general arguments in support of the instability at high $H\\,$. The first is related to causality. If, as in the standard $4$ dimensional case, the Hubble length $H^{-1}$ has the meaning of a casual horizon, we should expect that any stabilization mechanism cannot be at work when $H$ becomes much greater than the inverse size of ${\\cal M}$ (since different ``edges'' of ${\\cal M}$ would then be casually disconnected). Such a behavior can be found in~\\cite{lpps}, where the cosmology of the Randall-Sundrum model~\\cite{RS1} was studied under the assumption of radion stabilization. It was shown in~\\cite{lpps} that radion stabilization cannot be imposed if the physical energy on the hidden brane is greater than $\\sim {\\rm TeV^2} \\, M_p^2 \\,$. The appearence of this intermediate scale is somewhat surprising, and indeed it was left unexplained in~\\cite{lpps}. However, it can be verified that it is precisely at this energy density that $H^{-1}$ becomes smaller than the distance between the two branes.~\\footnote{We thank Lorenzo Sorbo for this observation.} A second argument can be inferred from the $dS_4 \\times S_d$ example. We noted that the instability has two possible end-points. One is characterized by a shrinking internal space, with asymptotic Kasner solution (a similar behavior was already noted in~(\\cite{branecode}) in the case of brane collisions). The second one is instead exact $dS_{4+d}\\,$, with all the compact and noncompact dimensions expanding with the same asymprtotic rate. This suggests that the instability can be attributed to the tendency of an inflationary expansion to homogenize and isotropize the {\\it whole} geometry, in this case by ``dragging'' also the compact space into a de--Sitter solution. This behavios, which is a well known feature of inflationary expansions, is encountered also in many Bianchi models. In the final part of the work, we have also commented on the possible role of the instability in the generation of primordial perturbations. The scalar perturbations appear as a conformal mode from the four dimensional point of view (for example, for an observer living on a $3$-brane, or after integrating out the internal space ${\\cal M}\\,$). This mode is a clear signature of the extra dimesnions, since it is absent in the standard $4$ dimensional case. If, due to the tachyonic contribution, this mode light ($0 < m < H$) during the de Sitter stage, it will be amplified with a scale invariant spectrum and with an amplitude which is disentangled from the standard result of slow-roll inflation. One can think of a number of ways how the conformal perturbation could be responsible for the adiabatic mode in the later FRW evolution. For instance, consider a mixture of fields conformally and nonconformally coupled to the noncompact geometry. Only the nonconformally coupled fields will be sensitive to the conformal perturbations, so that we expect that the final mode perturbation will be proportional to the amount of these fields in the mixture. In the text we have instead focused on a much simpler possibility, related to the mechanism of modulated perturbations~\\cite{modulated,dgz}. The conformal factor can be usually rescaled away by a rescaling of the energy scales in the $4$ dimensional theory (this is exactly what occurs in the Randall-Sundrum model~\\cite{RS1}). Hence, its fluctuations can be interpreted as fluctuations masses of particles and rates of processes in the $4$ dimensional theory. In oarticular, this may be the origin of the fluctuations of the decay rate of the inflaton, which is at the basis of the mechanism of modulated perturbations." }, "0403/astro-ph0403647_arXiv.txt": { "abstract": "We describe measurements of the mirror vignetting in the XMM-Newton Observatory made in-orbit, using observations of SNR G21.5-09 and SNR 3C58 with the EPIC imaging cameras. The instrument features that complicate these measurements are briefly described. We show the spatial and energy dependences of measured vignetting, outlining assumptions made in deriving the eventual agreement between simulation and measurement. Alternate methods to confirm these are described, including an assessment of source elongation with off-axis angle, the surface brightness distribution of the diffuse X-ray background, and the consistency of Coma cluster emission at different position angles. A synthesis of these measurements leads to a change in the XMM calibration data base, for the optical axis of two of the three telescopes, by in excess of 1 arcminute. This has a small but measureable effect on the assumed spectral responses of the cameras for on-axis targets. ", "introduction": "\\label{sect:intro} % XMM-Newton \\cite{Jansen} comprises 3 co-aligned telescopes, each with effective area at 1.5keV of $\\sim$1500cm$^{2}$, and Full Width Half Maximum (FWHM) angular resolution of $\\sim$5 arcseconds. The 3 telescopes each have a focal plane CCD imaging spectrometer camera provided by the EPIC consortium. Two also have a reflection grating array, which splits off half the light, to provide simultaneous high resolution dispersive spectra. These two telescopes are equipped with EPIC MOS cameras \\cite{mos}, which are conventional CMOS CCD-based images enhanced for X-ray sensitivity. The third employs the EPIC PN camera \\cite{pn} which is based on a pn-junction multi-linear readout CCD. The EPIC cameras offer a field of view (FOV) of $\\sim$30 arcminute diameter, and an energy resolution of typically 100~eV (FWHM) in the range $\\sim$0.2--10~keV. The two MOS telescopes are equipped with a Reflection Grating instrument \\cite{denHerder} that has its own dedicated readout camera. The in-orbit calibration of the XMM-Newton mirrors has been reported elsewhere \\cite{Aschenbach}, with special reference to the on-axis angular resolution (Point Spread Function, PSF). A second important calibration data set that is critical for analyzing spectroscopic information is the energy-dependent effective area \\cite{Aschenbach2002}. Both these features are under constant review as a result of improving knowledge of the instrumentation, and the requirements imposed by new science investigations. In this work we concentrate on different aspects of mirror performance that must be calibrated in the context of other scientific drivers which include, for example, cluster radial brightness distribution for determining gas mass, exposure maps and counts-to-flux conversions in population studies and diffuse background normalization measurements etc.. The reduction in effective area with radial distance from the field of view centre, or vignetting, must be accurately determined to support these investigations. To highlight the effect visually, Figures~\\ref{seren_image} and \\ref{seren_image2} show the excess flux per source detected in the 1XMM catalogue of EPIC serendipitous source detections (plotted in units of sigma). The images are displayed in the EPIC camera detector coordinates, and flux determinations assume the nominal vignetting correction centred on the reference pixel of the detector co-ordinate system (DETX, DETY in the nomenclature of the XMM data analysis system). These figures show that some low level discrepancy in the spatial variation in effective area calibration must be present. \\begin{figure} \\caption{Image in the MOS-2 detector plane, of the mean difference in the total-band (0.2-12 keV) flux seen by MOS-2 and MOS-1 expressed in Sigma. Bright pixels indicate an excess of flux in MOS-2 and dark pixels an excess in MOS-1. } \\label{seren_image} \\begin{tabular}{c} \\includegraphics[bb = 20 100 573 740, clip,scale=0.65]{m2flux2_box.ps} \\end{tabular} \\end{figure} \\begin{figure} \\caption{ The equivalent image for the Epic-pn detector plane, of the mean difference in the total-band (0.2-12 keV) flux seen by Epic-pn and MOS-1 expressed in Sigma. Bright pixels indicate an excess of flux in pn and dark pixels an excess in MOS-1. } \\label{seren_image2} \\begin{tabular}{c} \\includegraphics[bb = 36 170 478 680, clip, scale=0.65]{pnflux2_box.ps} \\end{tabular} \\end{figure} For XMM-Newton, direct measurement on the ground of the X-ray vignetting function was prevented because nearly all X-ray beam measurements were performed in a non-parallel beam. The installation of an X-ray stray-light baffle in front of the mirrors, and the Reflection Grating Array (RGA) stack at the mirror exit plane \\cite{denHerder}, introduced potential complications that were only measured in long wavelength, visible light at an EUV parallel beam facility \\cite{Tock}. Although the measured {\\em geometric} vignetting factor at longer wavelengths was comparable with predictions, it was necessary to use in-orbit data to confirm the X-ray energy dependence, and check that the geometric factor was maintained through the spacecraft assembly, integration, verification and launch campaigns. ", "conclusions": "The energy dependent vignetting calibration can be well matched to pre-launch predictions, but only on an assumption that the telescope optical axis is not perfectly aligned with the telescope boresight. This is not unexpected following difficulties on-ground of maintaining and/or measuring the telescope axis to better than 10's arcseconds. We note that the assumed telescope axis misalignment implies that ``on-axis'' targets at the common boresight location are actually at a slightly different vignetting value per telescope. We speculate that this partly accounts for some of the observed flux discrepancies between the MOS and pn cameras (Figure~\\ref{vigcal}). After reviewing these data sets, it was decided that the XMM-Newton calibration database would be updated in 2004 to account for new reference axes to be centred at PN (DETX,DETY = 1240,400 ), MOS1 (DETX,DETY = 100,-200 ) and MOS2 (DETX,DETY = 500,-1250 ). \\begin{figure}[ht] \\begin{center} \\begin{tabular}{c} \\includegraphics[scale=0.5]{vigcal.ps} \\end{tabular} \\end{center} \\caption{ Ratio of uncorrected and vignetted effective areas for an on-axis target. The flux differences for typical 0.5-2keV band will be $\\sim$ a few \\% for typical objects, and a small difference in recovered spectral slope will be caused by the change with energy }\\label{vigcal} \\end{figure}" }, "0403/astro-ph0403192_arXiv.txt": { "abstract": "We present observations of a remarkable submillimetre-selected galaxy, SMM\\,J16359+6612. This distant galaxy lies behind the core of a massive cluster of galaxies, A\\,2218, and is gravitationally lensed by the foreground cluster into three discrete images which were identified in deep submillimetre maps of the cluster core at both 450 and 850$\\mu$m. Subsequent follow-up using deep optical and near-infrared images identify a faint counterpart to each of the three images, with similar red optical--near-infrared colours and {\\it HST} morphologies. By exploiting a detailed mass model for the cluster lens we estimate that the combined images of this galaxy are magnified by a factor of $\\sim$45, implying that this galaxy would have un-lensed magnitudes $K_s=22.9$ and $I=26.1$, and an un-lensed 850-$\\mu$um flux density of only 0.8\\,mJy. Moreover, the highly constrained lens model predicted the redshift of SMM\\,J16359+6612 to be $z=2.6\\pm0.4$. We confirm this estimate using deep optical and near-infrared Keck spectroscopy, measuring a redshift of $z=2.516$. SMM\\,J16359+6612 is the faintest submm-selected galaxy so far identified with a precise redshift. Thanks to the large gravitational magnification of this source, we identify three sub-components in this submm galaxy, which are also seen in the NIRSPEC data, arguing for either a strong dust (lane) absorption or a merger. Interestingly, there are two other highly-amplified galaxies at almost identical redshifts in this field (although neither is a strong submm emitter). The three galaxies lie within a $\\sim 100$\\,kpc region on the background sky, suggesting this submm galaxy is located in a dense high-redshift group. ", "introduction": "Recent submillimetre (submm) surveys show that the majority of the submm background at wavelengths of $\\sim 850 \\mu$m arises from a population of distant, highly luminous infrared galaxies with 850\\,$\\mu$m flux densities of $\\gs 1$\\,mJy (Blain et al.\\ 1999; Cowie et al.\\ 2002; Smail et al.\\ 2002). The steep slope of the submm counts indicates that the integrated background is dominated by $\\sim 1$\\,mJy galaxies, below the $\\sim 2$\\,mJy confusion limit of the deepest blank-field 850$\\mu$m surveys carried out to date (Hughes et al.\\ 1998), and well below the 3--8\\,mJy flux density limits (3--4$\\sigma$) typically achieved in wider-field 850\\,$\\mu$m surveys (e.g.\\ Eales et al.\\ 1999; Scott et al.\\ 2002; Webb et al.\\ 2003). Submm galaxies in the 1-mJy regime can, however, only be studied at present by exploiting a natural telescope -- a massive gravitational lens formed by a rich clusters of galaxies -- which provide the opportunity of overcoming both the confusion limit (by spatially magnifying an area of the sky behind the lens) and the sensitivity limit (by gravitational magnification of the sources within this area). This strategy has successfully been used by Smail, Ivison \\& Blain (1997), Chapman et al.\\ (2002), Cowie et al.\\ (2002) and Smail et al.\\ (2002) to detect submm sources amplified by factors typically of 1.5--4$\\times$. Unfortunately, the intrinsically faintest sources are usually only detected at modest significance, complicating the identification and analysis of their counterparts in other wavebands. To investigate the properties of the mJy-population in detail we require a high-signal-to-noise detection of a sub-mJy submm galaxy. Such systems can arise in rare configurations where the source is located within a small region on the background sky defined by the caustic of the lens, which then forms multiple, highly magnified images of the background galaxy (see also Borys et al.\\ 2004). This situation not only provides a tool for investigating the properties of intrinsically faint galaxies in detail, but also provides a unique opportunity to measure the redshift of the background galaxy from the geometry of the images combined with an accurate model of the cluster lens, using a ray-tracing technique to triangulate the three-dimensional position of the galaxy in the volume behind the lens. This technique, which was demonstrated successfully by Kneib et al.\\ (1996) and confirmed by Ebbels et al.\\ (1998), circumvents the need for bright optical or infrared counterparts to a submm galaxy prior to measuring its redshift. We may thus avoid a bias in redshift measurements towards unrepresentative, low redshift or less obscured systems (Smail et al.\\ 2002; Webb et al.\\ 2003). This paper describes the discovery of a multiply-imaged submm galaxy, SMM\\,J16359+6612, which is gravitationally lensed by the core of the rich cluster A\\,2218, and appears as three distinct sources in our 850\\,$\\mu$m discovery map. We identify faint counterparts to all three sources in deep optical and near-infrared (NIR) imaging, consistent with their identification as three images of a single background galaxy. We estimate its redshift using a detailed mass model for the cluster lens. We then confirm the accuracy of the redshift with spectroscopic observations and finally discuss the intrinsic properties of this system. Throughout we will assume an $\\Omega=0.3$, $\\Lambda=0.7$ cosmology with $H_0=70$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$. At a redshift of $z=2.516$ the angular scale is thus 8.06 kpc/arcsec. ", "conclusions": "\\label{sec.discussion} The occurrence of a submm galaxy falling within the caustic lines of a massive foreground gravitational lens provides unique constraints on the properties of this faint submm galaxy. Using the cluster mass model we estimate that the background galaxy is gravitationally amplified by a factor of $\\sim$45 (integrated across all three images), indicating that the intrinsic 850\\,$\\mu$m flux density of this galaxy would be 0.8\\,mJy in the absence of gravitational magnification, while in the optical and NIR the galaxy would have magnitudes $I_{814}=26.1$ and $K_s=22.9$. The galaxy therefore represents a serendipitously positioned example of the submm galaxy population at flux levels of $\\sim 1$\\,mJy, the population which produces the bulk of the submm background (Blain et al.\\ 1999). This provides a unique opportunity to compare the properties of this low-luminosity submm galaxy with those of more luminous, submm galaxies studied in brighter, blank-field surveys (e.g.\\ Ivison et al.\\ 2002; Chapman et al.\\ 2003). A second, multiply-imaged submm galaxy has been identified by Borys et al.\\ (2004), giving two high-magnification examples from the 24 cluster lenses at $z>0.1$ mapped with sufficient sensitivity using SCUBA (Chapman et al.\\ 2002; Smail et al.\\ 2002; Knudsen et al.\\ in prep.). The surface density of $\\geq 1$\\,mJy submm galaxies is around 3\\,arcmin$^{-2}$ (Smail et al.\\ 2002; Cowie et al.\\ 2002), suggesting that roughly ten clusters need to be surveyed to detect a strongly-lensed mJy-flux submm galaxy. This is consistent with the expectation based on models for the cluster lenses of uniform $\\sim$800\\,km\\,s$^{-1}$ potential wells at $z\\sim 0.2$, each with a critical curve encircling 0.05\\,arcmin$^{-2}$, and background submm galaxies at $z\\sim 2.4$ (see Kraiberg et al 2004 for a thorough discussion). In our assumed cosmology the far-infrared luminosity of SMM\\,J16359+6612 is $1.0\\times10^{12}$\\,L$_\\odot$, roughly 1.5$\\times$ fainter than Arp\\,220, close to the border-line between luminous infrared galaxies and ultraluminous infrared galaxies (ULIRGs). The star formation rate derived from the far-infrared luminosity is about 500\\,M$_\\odot$\\,yr$^{-1}$. Figure~3 compares the restframe (lens-corrected) SED of the SMM\\,J16359+6612 to the SEDs of Arp\\,220, the well-studied $z=2.8$ submm galaxy SMM\\,J02399$-$0136 and the $z=1.44$ Extremely Red Object and submm galaxy HR10 (Dey et al.\\ 1999). The L$_{FIR}$/L$_{opt}$ ratio is higher than that seen for Arp\\,220, suggesting it is more obscured, and similar to the high-redshift, but much more luminous, SMM\\,J02399$-$0136 (Ivison et al.\\ 1998). \\smallskip \\centerline{\\psfig{file=a2218_multi_sed_col5.ps,width=3.4in}} {\\small\\addtolength{\\baselineskip}{-1pt} \\noindent{\\sc Fig.~4.} --- The observed Spectral Energy Distribution (SED) of the multiply-imaged submm galaxy SMMJ\\,16359+6612 (solid points with error bars) compared to Arp\\,220 (solid line, composite spectrum from Anantharamaiah et al.\\ 2000, Klaas et al.\\ 1997, Lisenfeld et al.\\ 1996 and Surace et al.\\ 2000) which is 1.5$\\times$ more luminous in the far-infrared, as well as the well-studied submm galaxy SMMJ\\,02399$-$0136 (open diamonds, at $z=2.8$, Ivison et al.\\ 1998) which is 10$\\times$ more luminous in the far-infrared, and the Extremely Red object HR10 (open triangles, at $z=1.44$, Dey et al.\\ 1999) which is 7$\\times$ more luminous in the far-infrared. All SEDs are redshifted to match that of SMM\\,J16359+6612 and scaled so they have a similar far-infrared luminosity. Notice the relatively large variation of L$_{FIR}$/L$_{opt}$ between the different galaxies.\\\\ \\label{fig.sed} } At $z=2.515$ the spatial resolution of the {\\it HST} images, corresponds to $\\sim 0.1$\\,kpc, taking into account the gravitational magnification. Thus the colour gradient within the background galaxy is on 1--2\\,kpc scales -- similar to the obscured region in nuclear starbursts in ultraluminous infrared galaxies at the present-day. A constraint on the size of the submm emission region is obtained by subtracting point sources from the 850 and 450\\,$\\mu$m maps. This procedure leaves no detectable residual of extended emission, indicating that to our measurement accuracy we are dealing with point sources. The tightest constraint is obtained at 450\\,$\\mu$m where we find that the intrinsic extent of the emission is less than $4''$, which corresponds to $\\sim 8$\\,kpc. There is thus no indication for a highly extended obscured starburst in this relatively low-luminosity submm galaxy. The precise redshift we have measured for SMM\\,J16359+6612 enables us to identify that this galaxy lies at an identical redshift to that of \\#273 and the \\#384/\\#468 multiple image system. Using the lens model we estimate that SMM\\,J16359+6612 lies between the \\#384/\\#468 and \\#273 galaxies in the source plane, and all three galaxies are less than 130\\,kpc apart. If these galaxies were not magnified by the cluster lens, all three galaxies would appear within a radius of 8$''$. Moreover, we can place a limit of $\\ls 100$\\,km\\,s$^{-1}$ on the possible velocity offset between these three galaxies. They are thus all part of a single group and it is likely that they are interacting, which may explain the activity we detect in the submm waveband. The strong clustering of the submm galaxies with UV-bright populations highlights the opportunity for measuring the distances to submm galaxies from the redshifts of less-obscured companions, as well as the possible confusion which may arise when trying to relate UV- and submm-selected populations, in the absence of precise positions from radio counterparts. SMM\\,J16359+6612 has much redder colour in $(V-I)$ and $(I-K)$ than the two other nearby UV-selected galaxies, confirming the dustier nature of this galaxy. If we compare the unlensed star formation rate (SFR) for this three systems, we find based on their UV continuum that: SMM\\,J16359+6612 has 6 M$_\\odot$\\,yr$^{-1}$ (although certainly underestimated due to dust extinction) \\#468: 14 M$_\\odot$\\,yr$^{-1}$ and \\#273: 4 M$_\\odot$\\,yr$^{-1}$. For SMM\\,J16359+6612 we can compare the three estimates of its amplification-corrected star formation rate: 6 M$_\\odot$\\,yr$^{-1}$ from its dust-corrected UV luminosity, 11\\,M$_\\odot$\\,yr$^{-1}$ from the H$\\alpha$ flux (uncorrected for extinction or aperture losses) and 500\\,M$_\\odot$\\,yr$^{-1}$ based on the far-infrared emission. Applying the median extinction correction derived for mid-IR selected luminous infrared galaxies at $z\\sim 0.7$ by Flores et al.\\ (2003), A$_{{\\rm H}\\alpha}\\sim 2.1$, and a modest correction for slit losses would increase the H$\\alpha$-derived star formation rate by a factor of $\\sim 10\\times$. These results suggest the vast majority of the young stars in this galaxy are obscured by dust and are undetectable in the restframe far-UV. Although it is 10$\\times$ less luminous than the typical submm-selected galaxy studied in blank-field surveys, this system shares the same extreme levels of obscuration seen in the more luminous galaxies, rather than the more modest dust obscuration inferred for the somewhat less luminous UV-selected galaxies at these redshifts. We also note that the H$\\alpha$ line width and velocity structure, if they reflect the dynamics of the galaxy, suggest that this system is more massive for its UV luminosity than typical UV-selected galaxies at this epoch (Erb et al.\\ 2003). In summary, we have identified a multiply-imaged submm galaxy seen through the core of the rich cluster A\\,2218. The cluster lens amplifies the background galaxy by a factor of $\\sim 45\\times$, providing a high signal-to-noise view of an example of the sub-mJy submm population which provides the bulk of the extragalactic background in this waveband. We estimate a redshift for this galaxy from our highly-constrained lens models for the cluster and confirm this using optical and near-infrared spectroscopy from Keck. The redshift of the submm galaxy is $z=2.516$, placing it at the same redshift as two other strongly-lensed UV-bright galaxies in this field. Our Keck spectroscopy suggests that the emission from the submm galaxy is dominated by star formation and in contrast to the typical UV spectral properties of more luminous submm galaxies, this galaxy shows Ly-$\\alpha$ absorption, rather than emission (c.f.\\ Chapman et al.\\ 2003). The star formation rates we derive from the UV continuum, H$\\alpha$ and far-infared emission show that most of the star formation activity is obscured and we suggest it is likely to be located in the component $\\gamma$, given its extreme $(I-K)$ colour. Because of its faint submm flux this galaxy is likely to be a good example of the type of galaxy that makes most contribution to the star formation history, thus deserving a detailled study with submm/mm interferometer to study the dynamics and mass of this obscured region. The presence of three highly-amplified $z=2.515$ galaxies in our survey field indicates that the submm galaxy resides in a compact group, interactions within which may help explaining the triggering of the obscured starburst we detect." }, "0403/astro-ph0403471_arXiv.txt": { "abstract": "A plausible model is proposed for the enhancement of the abundance of molecular species in bipolar outflow sources. In this model, levels of HCO$^+$ enhancement are considered based on previous chemical calculations, that are assumed to result from shock-induced desorption and photoprocessing of dust grain ice mantles in the boundary layer between the outflow jet and the surrounding envelope. A radiative transfer simulation that incorporates chemical variations within the flow shows that the proposed abundance enhancements in the boundary layer are capable of reproducing the observed characteristics of the outflow seen in HCO$^+$ emission in the star forming core L1527. The radiative transfer simulation also shows that the emission lines from the enhanced molecular species that trace the boundary layer of the outflow exhibit complex line profiles indicating that detailed spatial maps of the line profiles are essential in any attempt to identify the kinematics of potential infall/outflow sources. This study is one of the first applications of a full three dimensional radiative transfer code which incorporates chemical variations within the source. ", "introduction": "The complex morphologies of molecular outflows have been identified in numerous studies, mostly based on high resolution, interferometric CO and optical surveys (eg. Arce \\& Goodman, 2001, 2002; Lee et al. 2000, 2002). These surveys show that the molecular distributions are generally composed of a large scale, poorly collimated low velocity outflow that essentially traces the interaction between a surrounding molecular cloud envelope and a collimated high velocity jet. These morphologies are hard to interpret unambiguously but often resemble limb-brightened shells or sheaths surrounding cavities. Observations at single-dish resolution suggest that the abundance of HCO$^+$ in star forming cores with bipolar outflows may be enhanced by a factor of up to 100--1000$\\times$ relative to dark-cloud values - for example, a fractional abundance of $4\\times 10^{-8}$ is implied in the case of the Class 0 source B335 (Rawlings, Taylor and Williams 2000). This, and similar observations, provided the basis for the models of Rawlings~et~al.\\@~(2000) in which the enhancement of HCO$^+$ originates in the turbulent mixing layer which is the interface between the high velocity outflow and the quiescent or infalling core gas which it is steadily eroding. The molecular enrichment is driven by the desorption of molecular-rich ice mantles, followed by photochemical processing by shock-generated radiation fields. Thus, CO and H$_2$O are released following mantle evaporation. The CO is then photodissociated and the C that is produced is photoionized by the shock-induced radiation field that is generated in the interface: \\[ {\\rm CO} + {\\rm h}\\nu \\longrightarrow {\\rm C} + {\\rm O} \\] \\[ {\\rm C} + {\\rm h}\\nu \\longrightarrow {\\rm C}^+ + {\\rm e}^- \\] The subsequent reactions of C$^+$ with H$_2$O lead to an enhancement of the HCO$^+$ abundance: \\[ {\\rm C}^+ + {\\rm H}_2{\\rm O} \\longrightarrow {\\rm HCO}^+ + {\\rm H} \\] This abundance enhancement is only temporary, and progresses as a wave that fans out from the interface into the core. Such behaviour is both qualitatively and quantitatively consistent with the scenario proposed by Velusamy and Langer (1998) for B5 IRS1 on the basis of their $^{12}$CO~(2-1) and C$^{18}$O~(2-1) observations; the data clearly show a parabolic outflow cavity that appears to be steadily widening, so that the opening angle is growing at a rate of 0.006$^\\circ$yr$^{-1}$. An extension of the chemistry to a larger species/reaction set by Viti, Natarajan and Williams (2002) confirmed this result and concluded that other molecular tracers, such as H$_2$CS, SO, SO$_2$ and CH$_3$OH could be similarly affected. At higher resolutions, aperture synthesis observations are capable of identifying the morphology of the interaction between the jets and their surroundings. Hogerheijde et al. (1998) made high resolution HCO$^+$ J=1$\\to$0 observations using the Owens Valley Millimetre Array (OVRO) towards a number of low mass YSOs. One of the most interesting results from that survey was the detection of compact emission associated with the walls of an outflow cavity in L1527 (Hogerheijde et al. 1998, figure 1). The $^{12}{\\rm CO} 3-2$ outflow lobes are well-developed, with lengths of $\\sim 0.12~\\rm pc$ and a large opening angle ($\\sim 90^{\\circ}$). The single dish observations of Hogerheijde et al (1997) show that the HCO$^+$ emission is confined to a smaller scale, $\\sim 0.015$ to 0.022~pc, in the center of the flow. The interferometric observations with higher angular resolution show that much of this HCO$^+$ emission originates in the boundary of the lobes of the CO outflow rather than in the quiescent core of the cloud surrounding the outflow. The enhanced HCO$^+$ emission exhibits a cross-shaped morphology which is, of course, undetectable at single-dish telescope resolution (e.g. see Hogerheijde et al. 1997, figure 3). The data is also consistent with an HCO$^+$ enhancement by a factor of about 10$\\times$ in the outflow sheaths. The region of apparent HCO$^+$ emission excess has a limited extent, $20-30\\arcsec\\sim 3000-4500~{\\rm AU}$, at an assumed distance of 150~pc, a factor of 10 smaller than the CO emission. If the HCO$^+$ were enhanced over the length of the outflow, then more extended HCO$^+$ emission would be expected in the single-dish observations even if the enhanced region were at a scale too large to be detected by the interferometer (the OVRO will not recover any emission larger than $\\sim 30\\arcsec$). The small extent of the HCO$^+$ emission implies that the enhancement is temporary. For example, if the HCO$^+$ molecules created in the boundary layer of the outflow close in to the star are not destroyed, then the molecules ought to be transported down the outflow and ought to be detected in the single dish observations. The extent of the HCO$^+$ provides a constraint on the timescale for decay. The position-velocity maps for this source (Hogerheijde et al. 1998, figure 7) suggest that the outflow velocity in the emission-enhanced mixing layer is only $\\sim 5~{\\rm km~s^{-1}}$. This implies a dynamical age for the HCO$^+$ emitting gas of just $\\sim$3100 yrs, assuming the outflow is close to the plane of the sky. We therefore require a mechanism to restrict the HCO$^+$ enhancement to this region with a timescale of just a few thousand years. As stated above, this source has an outflow with a wide opening angle ($\\sim 90^{\\circ}$). Note that the $^{12}\\rm CO$ emission is much more extensive and has a larger dynamical age than the HCO$^+$ emission excess gas. The orientation of the outflow axis is close to the plane of the sky for this source which implies that the outflow and core emission will be not be clearly separable. It should also be noted that the dense core may be comparable to this size which could truncate the HCO$^+$ emission. However, there is still some ambiguity as to the interpretation of the morphology; whether it genuinely traces the abundance, temperature or density enhancements in the outflow interface, or whether it is a simply an excitation effect; the elevation of the HCO$^+$ excitation temperature in the boundary layer perhaps deriving from the outflow cavity acting as a low opacity pathway for photons from the star-disk boundary layer (Spaans et al. 1995). However, we must also recognise that it would be most unlikely for the HCO$^+$ abundance in the mixing layer to be the same as in the surrounding core. Previous models of boundary layers, including both analytical studies of turbulent interfaces (e.g. Rawlings and Hartquist 1997) and numerical hydrodynamical calculations (Lim, Rawlings and Williams 2001) have all shown strong HCO$^+$ abundance enhancements within the interface. Moreover, these studies have not included the effects of gas-grain interactions which are likely to further enhance the HCO$^+$ abundance. In this paper we do not attempt to model the chemical processes within the interface in any detail. Rather we consider the possible levels of molecular enhancement that are consistent with the observations. This is a timely precursor to the more detailed chemical/radiative transfer calculations that will be required as more higher resolution observational facilities become available. These facilities will be capable of resolving the dynamics and internal structures of sources on milliarcsecond scales. We have considered two different scenarios that can result in molecular enhancement (and specifically, of HCO$^+$) in the outflow sheath which forms the interface between the bipolar outflow jet and the molecular core envelope: \\begin{itemize} \\item Molecular enrichment through entrainment of dense circumstellar (disk) material into the outflow, and \\item Shock-induced mantle sublimation and photochemical enhancements (as in Rawlings~et~al.\\@~2000), dependent on the precise geometry of the outflow. \\end{itemize} ", "conclusions": "In this study we have concentrated on the implications of the radiative transfer modelling for our understanding of the chemical structure of bipolar outflow jet/boundary layers. We have not attempted to present or test a dynamical model of any sophistication. This is in contrast with some previous studies (e.g. Hogerheijde, 1998; Lee et al., 2000; Arce and Goodman, 2002) which have considered the hydrodynamic activity in the jet, boundary layer and surrounding cloud in some detail. Instead, although we do not constrain the hydrodynamics of the flow, our study makes some important conclusions concerning the origin and level of anomalous chemical activity within boundary layers, and how - using a state-of-the-art radiative transfer code - high resolution observations can be used to diagnose that activity. In his study, Hogerheijde (1998) utilized an axisymmetric non-LTE Monte-Carlo to study a boundary layer whose thickness is 1/5 of the total width of the (evacuated) jet cavity. His model of the boundary layer was based on a plausible Couette flow in which the density, flow velocity and temperature vary across the boundary layer in simple monotonic fashions, maintaining pressure balance at all points (Stahler, 1994). Whilst somewhat more sophisticated than our uniform density flow, the model had problems in reproducing the observed morphology of L1527 - and in particular the density contrast between the limb-brightened edges and the body of the outflow cavity. Indeed, reasonable fits to the observed morphology - the line intensities and spatial contrasts in L1527 - were only obtained if the HCO$^+$ abundance in the boundary layer is {\\em not} enhanced over that in the envelope, but the fractional ionization in the boundary layer is very high [X(e$^-$)$\\sim 10^{-4}$]. \\begin{figure*} \\psfig{file=MD740f5.eps,width=450pt,bbllx=0pt,bblly=0pt,bburx=581pt,bbury=544pt} \\caption{ Line profiles of HCO$^+~(1-0)$ for the best model. For clarity, a subset of the 96x96 array of line profiles is displayed: only the central regions are shown and only every fourth profile, in both X and Y directions, is displayed. The viewing angle is $0\\deg$. The tiny peaks of emission are from the cold envelope gas} \\label{plot5} \\end{figure*} \\begin{figure*} \\psfig{file=MD740f6.eps,width=450pt,bbllx=0pt,bblly=0pt,bburx=616pt,bbury=542pt} \\caption{Line profiles of HCO$^+~(1-0)$ for the best model. For clarity, a subset of the 96x96 array of line profiles is displayed: only the central regions are shown and only every fourth profile, in both X and Y directions, is displayed. The viewing angle is $60\\deg$. The tiny peaks of emission are from the cold envelope gas} \\label{plot6} \\end{figure*} We do not reproduce this result, and in particular we do not find that by raising the HCO$^+$ abundance in the boundary layer that the cavity `fills' due to increased opacity. What we do find is that the density and contrast are more sensitive to the adopted HCO$^+$ abundance enhancement and the thickness of the boundary layer, whilst the detailed morphology is dependent on the assumed outflow shape. In any case we do not find that such an extreme level of ionization is necessary. Indeed, it is not clear what the source of such a high level of ionization would be. More importantly, it must be remembered that the dominant loss route of HCO$^+$ in dark clouds is dissociative recombination, so it would seem reasonable to expect a significant {\\em suppression} of HCO$^+$ if X(e$^-$)$\\sim 10^{-5}-10^{-4}$. Our study is more restricted to a chemical analysis of the interface and thus adopts a much simpler dynamical model. Our modelling has attempted to reproduce the cross-shaped emission seen in L1527 (and other sources) with the appropriate intensity contrasts in the outflow lobes. The cause of the brightness in the cross-arms is essentially due to limb-brightening effects. Clearly, if the boundary layer is too thick then significant emission will be apparent along all lines of sight which pass through the boundary layer, and the morphology would tend to look more like a `bow tie' rather than a hollow cross. Our {\\em tanh} geometry is somewhat arbitrary, but was chosen because it matches well the observed morphology. The level of HCO$^+$ enhancement and our plausible explanation of its correlation with the shape of the outflow then yields patterns of intensity that closely resemble the high resolution maps; in particular the strong enhancements (and the location of the peaks of emission) in the limb-brightened wings, a natural explanation for the limited extent of the emission and the lack of emission close to the origin. The essential findings of this paper therefore are: \\begin{enumerate} \\item A strong enhancement of the HCO$^+$ abundance is required in the boundary layer between the outflow jet and the surrounding molecular core in order to explain the observations of certain outflow sources associated with star-forming cores \\item We have proposed a plausible, if simple, mechanism by which this enhancement can occur with the observed morphologies and contrasts; shock liberation and photoprocessing of molecular material stored in icy mantles. The degree of shock activity is closely related to the morphology of the source \\item We have shown how asymmetric, double-peaked line profiles can be generated with strong spatial variations in the relative strength of the red and blue wings. \\end{enumerate}" }, "0403/astro-ph0403701_arXiv.txt": { "abstract": "We have obtained moderate resolution ($\\sim 6$ km s$^{-1}$) spectroscopy of several hundred M giant candidates selected from Two Micron All Sky Survey photometry. Radial velocities are presented for stars mainly in the southern Galactic hemisphere, and the primary targets have Galactic positions consistent with association to the tidal tail system of the Sagittarius (Sgr) dwarf galaxy. M giant stars selected from the apparent trailing debris arm of Sgr have velocities showing a clear trend with orbital longitude, as expected from models of the orbit and destruction of Sgr. A minimum 8 kpc width of the trailing stream about the Sgr orbital midplane is implied by verified radial velocity members. The coldness of this stream ($\\sigma_v \\sim 10$ km s$^{-1}$) provides upper limits on the combined contributions of stream heating by a lumpy Galactic halo and the intrinsic dispersion of released stars, which is a function of the Sgr core mass. We find that the Sgr trailing arm is consistent with a Galactic halo that contains one dominant, LMC-like lump, however some lumpier halos are not ruled out. An upper limit to the total mass-to-light ratio of the Sgr core is 21 in solar units. Evidence for other velocity structures is found among the more distant ($>13$ kpc) M giants. A second structure that roughly mimics expectations for wrapped, leading Sgr arm debris crosses the trailing arm in the Southern Hemisphere; however, this may also be an unrelated tidal feature. Among the bright, nearby ($< 13$ kpc) M giants toward the South Galactic Pole are a number with large velocities that identify them as halo stars; these too may too trace halo substructure, perhaps part of the Sgr leading arm near the Sun. The positions and velocities of Southern Hemisphere M giants are compared with those of Southern Hemisphere globular clusters potentially stripped from the Sgr system. Support for association of the globular clusters Pal 2 and Pal 12 with Sgr debris is found based on positional and radial velocity matches. Our discussion includes description of a masked-filtered cross-correlation methodology that achieves better than 1/20 of a resolution element velocities in moderate resolution spectra. The improved velocity resolution achieved allows tighter constraints to be placed on the coldness of the Sgr stream than previously established. ", "introduction": "The extensive length of the tidal tails of the disrupting Sagittarius (Sgr) dSph system has recently been demonstrated in all-sky views of this system provided by the Two Micron All Sky Survey (2MASS) database (Majewski et al. 2003, ``Paper I\" hereafter). The relatively metal-rich stellar populations in the Sgr system means that M giant stars are prevalent in the Sgr debris stream. These stars are easily identified to distances of more than 60 kpc using 2MASS $JHK_s$ photometry, and primary leading and trailing tidal arms from Sgr are evident in all-sky M giant maps. However, as described in Paper I, while Sgr debris dominates the M giant population in the high halo, some ambiguities in the precise length and placement of the Sgr arms remain due to contamination by other M giant populations, particularly near the Galactic plane, and residual photometric errors. Checks on the membership of M giants to the streams using radial velocities can help delineate the extended morphology of the Sgr tidal tails. Stellar velocities are also useful tracers of mass. As well as revealing clues to the structure and integrity of the Sgr core itself, the placement and motions of its expansive tidal arms can provide important constraints on the mass of the Milky Way and the shape of its potential (e.g., Murai \\& Fujimoto 1980, Lin \\& Lynden-Bell 1982, Kuhn 1993, Johnston et al. 1999, Murali \\& Dubinski 1999, Ibata et al. 2001, Law, Johnston \\& Majewski 2004a,b). In addition, in principle, the velocity dispersion of tidal debris should provide a sensitive probe of the lumpiness of the halo (Moore et al. 1999, Johnston, Spergel \\& Haydn 2002, Ibata et al. 2002, Mayer et al. 2002). The present study represents a first effort to accumulate velocity data on 2MASS-selected M giants in the apparent Sgr debris trails. It includes a new radial velocity cross-correlation methodology that achieves better than 1/20 of a resolution element discrimination; this analysis approach, in combination with the fact that M giants are intrinsically bright and accessible with modest telescope apertures, places new, tighter constraints on the coldness of the Sgr stream using observations obtained with only a 1-m aperture telescope. We present results from a pilot radial velocity survey of 284 M giants from Paper I using the Swope 1-m telescope at Las Campanas Observatory. Most of the stars are in the Southern Galactic Hemisphere and lie within 5 kpc of the best-fit Sgr orbital plane (Paper I). Forty percent of the stars were selected to be in the distinct, Sgr Southern Arc (trailing tail), and, as we show below, most of these stars have velocities consistent with that association. The remaining stars are primarily very bright, nearby M giants toward the South Galactic Pole, which were selected for study to probe the possibility that Sgr tidal debris may be quite close to the Sun. We show that this latter sample likely contains an admixture of M giants from the Milky Way's Intermediate Population II/thick disk as well as stars from other halo substructure near the Sun --- perhaps nearby Sgr stars. Future papers will detail observations obtained at other telescopes subsequent to the data collected and analyzed here. ", "conclusions": "Radial velocities have been obtained for nearly three hundred M giant candidates identified in the 2MASS survey to be near the Sgr orbital plane. The survey data here concentrates mainly on stars in the Southern Hemisphere. The radial velocity trend of the Sgr trailing arm is clear and distinct and provides an important constraint on models of the Sgr debris stream (see Law et al. 2004a,b). The velocity dispersion of the trailing debris arm limits both the degree of lumpiness of the MW halo as well as the mass of the Sgr bound core. Because these two contributors to the velocity dispersion play off one another, it may be difficult simultaneously to have both a large $M/L$ for the Sgr core {\\it and} a lumpy halo: At present we limit the $(M/L)_{V,tot}$ of the Sgr core to 21 in solar units (see Law et al. 2004a,b), and the heating of the Sgr stream to be consistent with expectations for debris encounters with only the LMC. More stringent constraints on both may be possible after dynamical study of the Sgr tidal boundary to determine the zero-age velocity dispersion of Sgr debris. At least one other debris stream may be present in the distant M giant sample, and its velocity trend is consistent with model (Law et al. 2004a,b) predictions for wrapped, leading arm material in the Southern Hemisphere, but it may also be an unrelated tidal stream. The globular cluster Pal 12 may be associated with the stars in this feature, as previously postulated by Dinescu et al. (2000), but we note that the radial velocity of this cluster lies within 1.5-2$\\sigma$ of the mean for Sgr trailing arm M giants at this same location in the Galaxy. Based on its radial velocity and positional match with Sgr M giants, we find compelling evidence that the globular cluster Pal 2 is associated with Sgr trailing arm debris. Among the closer ($< 13$ kpc) M giant stars observed are more than half a dozen with the very high positive radial velocities expected for Sgr debris in the neighborhood of the Sun, and we point to the presence of many high velocity ($|v_{GSR}| > 150$ km s$^{-1}$), high latitude M giants that are difficult to account for as other than halo substructure relatively close to the Sun. The data presented here are better understood within the Sgr debris context by reference to interpretive models. Such models are given in Law et al. (2004a,b), in which the data presented here provide useful observational constraints. Moreover, additional velocities obtained in the Northern Hemisphere will help clarify the interpretations offered here. Future papers will present new data that increase by more than five times the present M giant velocity sample. We thank Carnegie Observatories Director A. Oemler for generous access to the Swope telescope and all of the Las Campanas staff for their diligent and generous support at the telescope. The results presented in this publication make use of data from the Two Micron All Sky Survey (2MASS), which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center (IPAC), funded by the National Aeronautics and Space Administration and the National Science Foundation. This work was supported by National Science Foundation grant AST-0307851 and a Space Interferometry Mission Key Project grant, NASA/JPL contract 1228235. MFS acknowledges support from NASA/JPL contract 1234021. MDW was supported in part by NSF grant AST-9988146. This work was also partially supported by the Celerity Foundation. We appreciate comments from the referee that have improved the clarity of the paper." }, "0403/astro-ph0403537_arXiv.txt": { "abstract": "The 692 GHz para ground-state line of D$_2$H$^+$ has been detected at the Caltech Submillimeter Observatory towards the pre-stellar core 16293E. The derived D$_2$H$^+$ abundance is comparable to that of H$_2$D$^+$, as determined by observations of the 372 GHz line of ortho-H$_2$D$^+$. This is an observational verification of recent theoretical predictions (Roberts, Herbst \\& Millar 2003), developed to explain the large deuteration ratios observed in cold, high-density regions of the interstellar medium associated with low mass pre-stellar cores and protostars. This detection confirms expectations that the multiply deuterated forms of H$_3$$^+$ were missing factors of earlier models. The inclusion of D$_2$H$^+$ and D$_3$$^+$ in the models leads to predictions of higher values of the D/H ratio in the gas phase. ", "introduction": "Recently, millimeter and submillimeter spectroscopy of the dense interstellar medium has shown that, in cold dense regions, deuterated molecular species are highly abundant, sometimes more than 10$^{-1}$ of the H version. Amazingly, doubly and triply deuterated species can be observed, e.g. D$_2$CO \\citep{ceccarelli98}, NHD$_2$ \\citep{roueff00}, CHD$_2$OH \\citep{parise02}, D$_2$S \\citep{vastel03}, ND$_3$ \\citep{lis02,vdtak02}, CD$_3$OH \\citep{parise04}. Several models have been developed to account for such high levels of deuteration \\citep{tielens83,roberts00a,roberts00b}. \\citet{phillips02} have pointed out that the deuteration of H$_3$$^+$ will be extended beyond H$_2$D$^+$, to D$_2$H$^+$ and D$_3$$^+$, and that detection of D$_2$H$^+$ might be possible. A calculation taking a high degree of deuteration into account has been carried out by \\citet{roberts03} and \\citet{walmsley04}, confirming the expectation that, in dense depleted regions, the abundance of D$_2$H$^+$ will be similar to that of H$_2$D$^+$, and that D$_3$$^+$ will be abundant. The key enabling work in the astronomical search for D$_2$H$^+$ is the laboratory measurement of the para ground-state transition (1$_{10}$-1$_{01}$) by \\citet{hirao03}. We report here the first astronomical detection of that transition.\\\\ Chemical reactions go in the direction to minimize energy. The chemical fractionation process favors the production of the heavier more deuterated species, because of the mass dependence of the zero-point vibration energies of the isotopic variants. Gas phase species are expected to be depleted at the centers of cold, dark clouds, since they accrete on the dust grains \\citep[see, e.g.,][]{charnley97}. A series of observations has shown that the abundances of molecules like CO decrease in many pre-stellar cores \\citep{bacmann02}. The removal of these reactive species affects the gas-phase chemistry and particularly the deuterium fractionation within the cloud. Indeed, the removal of species that would normally destroy H$_3$$^+$ \\citep[e.g. CO;][]{roberts00a} means that H$_3$$^+$ is more likely to react with HD and produce H$_2$D$^+$. For example, if [CO/H$_2$]~$\\sim$~5 $\\times$ 10$^{-6}$ \\citep{bacmann02}, this leaves HD at [HD/H$_2$]~$\\sim$~5 $\\times$ 10$^{-5}$ as the most abundant molecule available for reaction with H$_3^+$ and H$_2$D$^+$, and favors the production of high deuterium content molecules: \\begin{equation} H_3^+ + HD \\longleftrightarrow H_2D^+ + H_2 + \\Delta E_a \\end{equation} \\begin{equation} H_2D^+ + HD \\longleftrightarrow D_2H^+ + H_2 + \\Delta E_b \\end{equation} \\begin{equation} D_2H^+ + HD \\longleftrightarrow D_3^+ + H_2 + \\Delta E_c \\end{equation} where $\\Delta$E$_a$, $\\Delta$E$_b$ and $\\Delta$E$_c$ are the released energies of the exothermic reactions. Using the zero-point energies computed by \\citet{ramanlal03}, and the energy of the first allowed rotational state of the H$_3^+$ molecule permitted by the Pauli exclusion principle ($\\sim$ 92 K), these values are: $\\Delta$E$_a$ = $\\sim$ 230 K, $\\Delta$E$_b$ = $\\sim$ 180 K and $\\Delta$E$_c$ = $\\sim$ 230 K. After a long frustrating search \\citep{phillips85,pagani92,vandishoeck92,boreiko93}, and with the advent of new submillimeter receivers, H$_2$D$^+$ was detected toward two young stellar objects, NGC 1333 IRAS 4A (Stark et al. 1999) and IRAS 16293-2422A (Stark et al. 2004), although with relatively low signal strength. The H$_2$D$^+$ search has now been extended to pre-stellar cores, and has been detected with relatively strong emission (Caselli et al. 2003; Caselli et al. 2004, {\\it in preparation}; Vastel et al. 2004, {\\it in preparation}) confirming that H$_2$D$^+$ is dramatically enhanced in a gas depleted of most molecules. The ammonia and DCO$^+$ emission around the proto-binary system IRAS16293-2422 \\citep{wootten87,mizuno90,lis02} does not only peak on IRAS16293-2422 itself but shows a second peak, about 90$^{\\prime\\prime}$ to the southeast, in a condensation called 16293E. \\citet{lis02} found that CO in this region is depleted by a factor of 7. It is known to be a region where deuterium fractionation is strong and was chosen to be searched for D$_2$H$^+$. ", "conclusions": "From the observed line strengths, given in column 3 of Table \\ref{table1}, we estimate the H$_2$D$^+$ and D$_2$H$^+$ column densities (see Table \\ref{table2}) for an excitation temperature T$_{ex}$ of 10 K, assuming a 25\\% calibration uncertainty (3 $\\sigma$). The column density is given by: \\begin{equation} N_{tot}=\\frac{8\\pi\\nu^3}{c^3}\\frac{Q(T_{ex})}{g_uA_{ul}}\\frac{e^{E_{u}/T_{ex}}}{e^{h\\nu/kT_{ex}}-1}\\int_{}^{}\\tau\\, dv \\end{equation} where Q(T$_{ex}$) is the partition function, Assuming LTE conditions, we can estimate the optical depth from the observed line intensity: \\begin{equation} T_{mb} = [J_{\\nu}(T_{ex})-J_{\\nu}(T_{bg})](1-e^{-\\tau}) \\end{equation} where $J_{\\nu}(T) = (h\\nu/k)/(e^{h\\nu/kT}-1)$ is the radiation temperature of a blackbody at a temperature T, and T$_{bg}$ is the cosmic background temperature of 2.7 K. In the case of the H$_2$D$^+$ transition, g$_u$ = 9, A$_{ul}$ = 1.04 10$^{-4}$ s$^{-1}$, E$_{ul}$ = 17.9 K; in the case of the D$_2$H$^+$ transition, g$_u$ = 9, A$_{ul}$ = 4.55 10$^{-4}$ s$^{-1}$, E$_{ul}$ = 33.2 K. The derived column densities depend on the assumed value of the excitation temperature. Using NH$_3$, \\citet{mizuno90} estimate the gas temperature to be 12 K. Using D$_2$CO line ratios, \\citet{loinard01} obtained a rotational temperature of 8-10 K. Thus we quote, in Table \\ref{table2}, the values obtained for an excitation temperature of 10 K. Figure \\ref{figure4} presents the evolution of the ortho-H$_2$D$^+$ and para-D$_2$H$^+$ column densities as well as the para-D$_2$H$^+$/ortho-H$_2$D$^+$ ratio, as a function of temperature between 9 and 15 K. Figure \\ref{figure4} and Table \\ref{table2} represent the case where the source emission is extended compared to the beam size. If the source emission is comparable to or smaller than the CSO beam size, the para-D$_2$H$^+$/ortho-H$_2$D$^+$ ratio is then increased by a factor of $\\sim$ 1.5 at the average excitation temperature of 10 K. At thermal equilibrium, the ortho to para (respectively para to ortho) concentration ratio for H$_2$D$^+$ (respectively D$_2$H$^+$) is equal to 9 $\\times$ exp(-86.4/T) (respectively 9/6 $\\times$ exp(-50.2/T)), so that at 10 K, this ratio would be $\\sim$ 2 $\\times$ 10$^{-3}$ (respectively $\\sim$ 10$^{-2}$). However, taking into account the limited rates of the spin allowed collisions with H$_2$, it is found that at these low temperatures, the ortho to para H$_2$D$^+$ concentration ratio is close to unity \\citep{gerlich02}. The para to ortho D$_2$H$^+$ ratio is estimated by \\citet{walmsley04} to be about one, for the same conditions. The para-D$_2$H$^+$/ortho-H$_2$D$^+$ ratio presented in Figure \\ref{figure4} should then approximately represent the actual D$_2$H$^+$/H$_2$D$^+$ ratio. The 1.3 mm dust continuum strength \\citep[see][]{lis02} is $\\sim$ 0.3 Jy in a 11$^{\\prime\\prime}$ beam and $\\sim$ 1.3 Jy in a 20$^{\\prime\\prime}$ beam corresponding to the angular resolution of the H$_2$D$^+$ and D$_2$H$^+$ data. Assuming a dust temperature of 12 K and a mass opacity coefficient of 0.005 cm$^2$~g$^{-1}$ (appropriate for pre-stellar cores), we derive an H$_2$ column density of $\\sim$ 5~10$^{23}$ cm$^{-2}$. We then derive the H$_2$D$^+$ and D$_2$H$^+$ abundances to range between $\\sim$ 10$^{-10}$ (at 10 K) and $\\sim$ 10$^{-11}$ (at 15 K) compatible with abundances found by \\citet{roberts03} for a cloud at 10 K and n(H$_2$)=3 10$^6$ cm$^{-3}$. The main result of this work is {\\it the detection of D$_2$H$^+$, with an abundance comparable to that of H$_2$D$^+$}. This is a remarkable verification of recent theoretical predictions, aimed at explaining the large deuteration ratios observed in low mass pre-stellar cores and protostars. Two models, \\citet{roberts03} and \\citet{walmsley04}, have recently considered the effect of including all possible deuterated isotopomers of H$_3^+$ in the chemical networks, as suggested by \\citet{phillips02}. \\citet{roberts03} studied the temporal evolution of a cold and dense cloud, and found that at late times ($\\sim$ 10$^4$ yr), when CO is severely depleted in the gas phase (more than a factor of 1000), the D$_2$H$^+$/H$_2$D$^+$ ratio reaches unity. \\citet{walmsley04} studied the evolution of gas, depleted in CO, as a function of gas density and grain size distribution. For densities larger than 10$^6$ cm$^{-3}$, they also found that D$_3$$^+$ can be the most abundant ion and that the D$_2$H$^+$/H$_2$D$^+$ ratio reaches unity (see their figure 2). Both models need an extreme CO depletion to account for such a ratio. As discussed in the Introduction, the measured CO depletion in 16293E is a factor 7 \\citep{lis02}, rather than the extreme CO depletion needed by the models. However, the CO was measured in a large region (31\") compared with that probed by the present observations: 20$^{\\prime\\prime}$ for the H$_2$D$^+$ and 11$^{\\prime\\prime}$ for the D$_2$H$^+$ observations. Also, regions containing CO will not contain much H$_2$D$^+$ and D$_2$H$^+$ and vice-versa, so the inevitable inhomogeneities in the region inhibit a clear result. In summary, after some years of inconclusive results for theoretical models in understanding the observed high deuteration ratios of doubly and triply deuterated molecules, the present observation seems to suggest that the basic process is now almost completely understood: the large deuteration is due to extreme CO depletion, and the factor that was previously missing in the models is the multiply deuterated forms of H$_3^+$. This is quite an achievement, and one remaining step will be to verify that the last prediction, a significant abundance of D$_3^+$, is also fulfilled. At present, the lines detected here are the only ones available for H$_2$D$^+$ and D$_2$H$^+$. Knowledge of the ortho to para ratios and abundances of H$_2$D$^+$ and D$_2$H$^+$ would be considerably improved if the ground state transitions of para-H$_2$D$^+$ (at 1370.15 GHz) and ortho-D$_2$H$^+$ (at 1476.60 GHz) were available. These lines could be detected from space telescopes such as Herschel with the Heterodyne Instrument for the Far-Infrared and also possibly from the stratospheric observatory SOFIA. However, D$_3$$^+$, like H$_3$$^+$ has no permanent dipole moment. Therefore this molecule probably can only be detected in absorption in the near-infrared." }, "0403/astro-ph0403270_arXiv.txt": { "abstract": "We reported in a previous paper the discovery of large-scale structure of Lyman $\\alpha$ emitters (LAEs) at $z=4.86\\pm 0.03$ with a projected size of $20 h_{70}^{-1}$ Mpc $\\times$ $50 h_{70}^{-1}$ Mpc in narrow-band data of a $25' \\times 45'$ area of the Subaru Deep Field ($\\Omega_0=0.3, \\lambda_0=0.7, H_0 = 70 h_{70}$ km s$^{-1}$ Mpc$^{-1}$). However, the surveyed area, which corresponds to $55 h_{70}^{-1}$ Mpc $\\times 100 h_{70}^{-1}$ Mpc, was not large enough that we can conclude that we are seeing a typical distribution of $z\\simeq 5$ LAEs. In this Letter, we report the results of follow-up imaging of the same sky area using a new narrow-band filter (NB704, $\\lambda_c=7046$\\AA\\ and FWHM$=100$\\AA) to detect LAEs at $z=4.79$, i.e., LAEs lying closer to us by $39 h_{70}^{-1}$ Mpc on average than the $z=4.86$ objects. We detect 51 LAEs at $z=4.79 \\pm 0.04$ down to ${\\rm NB704}=25.7$, and find that their sky distribution is quite different from the $z=4.86$ LAEs'. The clustering of $z=4.79$ LAEs is very weak on any scales and there is no large-scale high-contrast structure. The shape and the amplitude of the angular correlation function are thus largely different between the two samples. These results demonstrate a large cosmic variance in the clustering properties of LAEs on scales of $\\sim 50 h_{70}^{-1}$ Mpc. ", "introduction": "\\label{sec:introduction} Search for Lyman $\\alpha$ emission of galaxies using a narrow-band filter is a powerful tool to detect high-$z$ faint galaxies. Indeed, many observations have successfully detected such Lyman $\\alpha$ emitters (LAEs) from $z\\sim2$ up to $z\\simeq 6.6$ (e.g., Hu, Cowie, \\& McMahon 1998; Pascarelle, Windhorst, \\& Keel 1998; Campos et al. 1999; Hu, McMahon, \\& Cowie 1999; Kudritzki et al. 2000; Rhoads et al. 2000; Stiavelli et al. 2001; Ajiki et al. 2002; Hu et al. 2002; Venemans et al. 2002; Fynbo et al. 2003; Kodaira et al. 2003; Ouchi et al. 2003a; Shimasaku et al. 2003; see also Maier et al. 2003). Narrow-band surveys can also map effectively large-scale distributions of LAEs (e.g., Steidel et al. 2000; Venemans et al. 2002; Hu et al. 2004). We recently made a survey of LAEs at $z=4.86$ using the narrow-band filter NB711 ($\\lambda_c=7126$\\AA, FWHM=73\\AA) in an area of $25' \\times 45'$, and found a large-scale structure of LAEs of $20h_{70}^{-1}$ Mpc $\\times$ $50h_{70}^{-1}$ Mpc size (Shimasaku et al. 2003). This is the first discovery of large-scale structure in young universes, suggesting that the birth of large-scale structure is very early in the history of the universe and that LAEs are strongly biased against dark matter, since Cold Dark Matter (CDM) models predict that the density fluctuations of dark matter are very small on such large scales. However, the size of the large-scale structure is nearly comparable to the size of the survey region ($55h_{70}^{-1}$ Mpc $\\times$ $100h_{70}^{-1}$ Mpc), and thus we cannot safely conclude that we are seeing a typical distribution of LAEs at $z\\simeq 5$. To address this issue, it is strongly needed to enlarge the survey volume. Motivated by this, we made a followup imaging survey of LAEs at $z=4.79$, i.e., LAEs located closer to us by $39h_{70}^{-1}$ Mpc than those at $z=4.86$, in exactly the same sky area. Using these data, we examine differences in the sky distribution of LAEs between the two redshifts. We adopt $\\Omega_0=0.3$, $\\lambda_0=0.7$, and $H_0 = 70 h_{70}$ km s$^{-1}$ Mpc$^{-1}$. ", "conclusions": "Hamana et al. (2004) found that the observed ACF of $z=4.86$ LAEs given in Ouchi et al. (2003a) cannot be reproduced by a simple halo model which assumes LAEs to be associated with dark haloes, because of too strong the observed correlation on scales $ \\gsim 2'$. We find that our $z=4.86$ LAEs have large-scale structure and that their ACF is high and not fit by a power law, while the clustering of $z=4.79$ LAEs is very weak. These results may suggest that the clustering of LAEs is typically weak, possibly tracing the dark-matter distribution, and that we happened to observe an unusual region in the $z=4.86$ universe where LAEs form a large, coherent structure of a size of $\\sim 50 h_{70}^{-1}$ Mpc. Conversely, if it turns out, from a larger survey, that the $z=4.86$ region we observed is relatively common in high-$z$ universes, this will suggest that LAEs and LBGs are separate populations in terms of clustering properties, since the clustering of LBGs has been found to be approximated well by halo models (Hamana et al. 2004). Detailed modeling of the clustering of LAEs based on an enlarged sample will give important hints on the nature of LAEs. Although we did not find in our two LAE samples a clear difference in the number density, a large field-to-field variance in the clustering of LAEs, including velocity structures, can influence measurements of the number density of LAEs based on narrow-band surveys, especially if the survey volumes are smaller than ours and if $b_g$ is much larger than unity; the $b_g$ value derived from the clustering of the $z=4.86$ LAEs is as large as $\\sim 10$. Shallower surveys will suffer from larger variances, since LAEs with brighter Lyman $\\alpha$ luminosities tend to be clustered more strongly (Ouchi et al. 2003a). For instance, Ajiki et al. (2003) found that the number density of $z=5.7$ LAEs in their sample of $2 \\times 10^5 h_{70}^{-3}$ Mpc$^3$ is three times higher than that estimated by Rhoads \\& Malhotra (2001) based on a similar survey volume. To summarize, our observations show that it is necessary to survey a much larger volume than ours in order to derive the average clustering properties of LAEs." }, "0403/astro-ph0403100_arXiv.txt": { "abstract": "s{ Recent \\chandra\\ and \\xmm\\ surveys have confirmed that the cosmic X-ray background is mostly due to accretion onto super-massive black holes, integrated over cosmic time. Here we review the results obtained from the photometric and spectroscopic follow-up observations of the 122 X-ray sources detected by the HELLAS2XMM 1dF Survey down to a 2--10~keV flux of $\\approx10^{-14}$~\\cgs. In particular, we focus on the multiwavelength properties of a few intriguing classes of X-ray sources: high X-ray-to-optical flux ratio sources, Type~2 quasars, and XBONGs.} ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403499_arXiv.txt": { "abstract": "We present FUSE spectroscopy and supporting data for star-forming regions in nearby galaxies, to examine their massive-star content and explore the use of abundance and population indicators in this spectral range for high-redshift galaxies. New far-ultraviolet spectra are shown for four bright H II regions in M33 (NGC 588, 592, 595, and 604), the H II region NGC 5461 in M101, and the starburst nucleus of NGC 7714, supplemented by the very-low-metallicity galaxy I Zw 18. In each case, we see strong Milky Way absorption systems from H$_2$, but intrinsic absorption within each galaxy is weak or undetectable, perhaps because of the ``UV bias\" in which reddened stars which lie behind molecular-rich areas are also heavily reddened. We see striking changes in the stellar-wind lines from these populations with metallicity, suggesting that C II, C III, C IV, N II, N III, and P V lines are potential tracers of stellar metallicity in star-forming galaxies. Three of these relations - involving N IV, C III, and P V - are nearly linear over the range from O/H=0.05--0.8 solar. The major difference in continuum shapes among these systems is that the giant H II complex NGC 604 has a stronger continuum shortward of 950 \\AA\\ than any other object in this sample. Small-number statistics would likely go in the other direction; we favor this as the result of a discrete star-forming event $\\approx 3$ Myr ago, as suggested by previous studies of its stellar population. ", "introduction": "The history of galaxies is, in large part, the history of star formation. Massive stars play key roles both as highly visible tracers of star formation and as players in altering surrounding star formation and both energy and chemistry of the interstellar medium. It is these stars which dominate the observed properties of actively star-forming galaxies. The massive part of stellar populations is most clearly observed in the ultraviolet, where their energy distributions peak and competing light from cooler stars is minimal, as long as the foreground extinction allows escape of enough of this radiation. In such environments, ultraviolet studies of star-forming regions have proven fruitful in understanding these populations. The recent opening of the far-ultraviolet window, between Lyman $\\alpha$ and the Lyman limit, for deep observations, allows study of massive hot stars in a range where they fully dominate the spectrum. This relatively narrow band contains an embarrassment of spectral riches, with numerous lines from stellar winds as well as interstellar material both atomic and molecular. These include the strong and highly-ionized lines of O VI and the unique ability to measure cold H$_2$. In addition, this piece of the spectrum is accessible for high-redshift galaxies, at least for composite samples where the Lyman $\\alpha$ forest can be averaged adequately, allowing the possibility of direct comparisons of stellar populations over a large span of cosmic time. The very sensitivity of the far-UV light to star formation and reddening makes it a purer probe of some properties of star-forming regions than observations at longer wavelengths. Since only short-lived stars contribute significantly in the far-UV range, the details of star-forming history should matter only for very brief bursts (such as might be found in individual H II regions, but are less likely on galaxy scales). This makes the far-UV spectrum more sensitive to the stellar population itself than to its history. Furthermore, although the extinction is high, its differential effect across the far-UV band is modest, and paradoxically the effective reddening to the stars we see is smaller than found at longer wavelengths. In observing stars intermingled with highly structured dust distributions, the ``picket-fence\" effect (Heisler \\& Ostriker 1988) means that most of the stars are so reddened as to make no significant contribution in the deep ultraviolet; all the stars we see are only lightly reddened. This also reduces the effects of the forest of H$_2$ absorption features because of the mixing of molecular gas and dust. Observational data on nearby galaxies, before the {\\it Far-Ultraviolet Spectroscopic Explorer} (FUSE), were limited to a handful of star-forming systems. Four starburst galaxies were observed using HUT on {\\it Astro-2} (Leitherer et al. 1995, 2002), largely to measure escaping radiation in the Lyman continuum, which provided initial data for comparison with models based on stellar spectra dating back to {\\it Copernicus}. The strongest features fall into two blends near 970 (Ly $\\gamma$ + C III) and 1030 \\AA\\ (O VI + Ly $\\beta$ + C II). A {\\it Voyager 2} observation of M33, with some spatial resolution in one direction, was analyzed by Keel (1998), showing that its far-UV continuum is virtually identical to those of the powerful starbursts, and that NGC 604 is bluer in this range than the overall disk. These data also suggested significant Lyman $\\alpha$ emission, with spatial profile suggesting an origin in the diffuse ISM rather than giant H II regions. Ironically, until the availability of FUSE observations, the richest information on galaxies shortward of Lyman $\\alpha$ came from objects at high redshift, particularly composite spectra of Lyman-break galaxies (Steidel et al. 2001, Shapley et al. 2003). To enable comparisons between local, well-studied star-forming systems and these powerful, young objects, we have undertaken a series of FUSE observations of star-forming regions innearby galaxies. We present here the analysis of these spectra in the context of their stellar populations and systematic changes with metallicity. A companion paper (Keel, Shapley, \\& Steidel 2004) considers the comparison with composite spectra of Lyman-break galaxies. ", "conclusions": "We have used FUSE spectra of star-forming regions in nearby galaxies, whose gas-phase metallicities range from 0.05-0.8 solar, to explore the utility of far-ultraviolet spectra in measuring the abundances in star-forming galaxies, as well as to probe the massive-star populations in these galaxies. The absorption lines from radiatively-driven winds prove to be very sensitive to metal abundance; all six strong and unblended species (including C IV from archival IUE data) have a strong, monotonic metallicity dependence. For N IV, C III, and P V, the relation between straighforward equivalent-width values and oxygen abundance from emission-line spectra is closely linear, suggesting that these lines will be useful in tracing the chemical history of galaxies from $z=3-4$, beyond which the Lyman $\\alpha$ forest makes even composite spectra progressively less informative. The continuum of NGC 604 departs from the uniform shape of the other objects below 950 \\AA . After considering the effects of small-number statistics among the massive stars in these objects, we conclude that this difference probably traces to a discrete burst of star formation $\\approx 3$ Myr ago in NGC 604. This region had been considered by several previous studies to have hoisted such a burst, on grounds of both morphology of the gas and fitting of the H-R diagram." }, "0403/astro-ph0403385_arXiv.txt": { "abstract": "{ Recently, combining radial velocities from Keck/HIRES \\'echelle spectra with published proper motion membership probabilities, \\citet{Coe_02} observed a sample of $21$ stars, probable members of Palomar~13, a globular cluster in the Galactic halo. Their projected velocity dispersion $\\sigma_\\mathrm{p} = 2.2 \\pm 0.4$~km\\thinspace s$^{-1}$\\ gives a mass-to-light ratio $\\mathcal{M}/L_V = 40^{+24}_{-17}$, about one order of magnitude larger than the usual estimate for globular clusters. We present here radial velocities measured from three different CCD frames of commissioning observations obtained with the new ESO/VLT instrument FLAMES (Fibre Large Array Multi Element Spectrograph). From these data, now publicly available, we measure the homogeneous radial velocities of eight probable members of this globular cluster. A new projected velocity dispersion $\\sigma_\\mathrm{p} = 0.6$--$0.9 \\pm 0.3$~km\\thinspace s$^{-1}$\\ implies Palomar~13 mass-to-light ratio $\\mathcal{M}/L_V = 3$--$7$, similar to the usual value for globular clusters. We discuss briefly the two most obvious reasons for the previous unusual mass-to-light ratio finding: binaries, now clearly detected, and more homogeneous data from the multi-fibre FLAMES spectrograph. ", "introduction": "All the dynamical studies of nearby globular clusters have established that these dynamical systems contain no dark matter, apart from the expected stellar remnants such as white dwarves and neutron stars \\citep[e.g.,][]{PrrMen93,MenHee97}. Consequently, globular clusters may be the most massive stellar systems in which no non-baryonic dark matter is dynamically detected, while dynamical evidence for non-baryonic dark matter seems to be present in most galaxies, from the faintest dwarf spheroidals (dSphs) to the brightest cDs galaxies, and clusters of galaxies as well. Some of the local dSphs, around the Galaxy and M31, have integrated absolute luminosities similar or fainter than those of the brightest Galactic globular clusters. Since there is evidence that some Galactic dSph galaxies are dark-matter-dominated, it is therefore reasonable to check if some globular clusters do present dynamical evidence for non-baryonic dark matter. Such a possibility may be emphasized by the current predictions of Cold Dark Matter (CDM) numerical simulations of galaxy formation, in which the number of low-mass dark-matter substructures orbiting the halo of massive galaxies largely exceeds the number of dwarf galaxies observed in the halos of both our Galaxy and M31 \\citep[e.g., ][]{Kln_99,Moe_99,Moe_01}. However, recently improved CDM models may correctly predict the observed number of satellite galaxies \\citep{Biy03}. Could some of the globular clusters in the outer parts of the Galactic halo be such dark-matter substructure~? These remote stellar systems have so far been poorly studied because of the difficulties in the acquisition of high-quality radial velocities and proper motions, direct consequences of the faintness and sparsity of these distant stellar systems. They are nevertheless important probes of the formation and evolution of the Galaxy, as their ages and metallicities provide direct constraints on the duration of halo formation process and on the time-scale for Galactic chemical enrichment, while the shape and extent of the Galactic dark halo are constrained by their orbital properties. In 1998, a program was started at Californian Institute of Technology to study the internal dynamics of seven distant halo globular clusters using the High Resolution Echelle Spectrometer (HIRES) at the W.~M.~Keck Observatory. The aim was the first direct measurements of the velocity dispersions and mass-to-light ratios for these clusters. Six clusters in this sample exhibited velocity dispersions $\\sigma_\\mathrm{p} \\sim 1$~km\\thinspace s$^{-1}$, translating into mass-to-light ratio values typical of globular clusters $\\mathcal{M}/L_V\\sim 3$ (all mass-to-light ratios quoted in this paper are in solar units). Only one cluster, the halo globular cluster Palomar~13, displayed a velocity dispersion larger than expected. \\citet[ also referenced below as the Keck study]{Coe_02} presented a careful analysis, combining radial velocities from Keck/HIRES \\'echelle spectra with published proper motion membership probabilities from \\citet{Sil_01}. They obtained a sample of 21 stars, probable members of Palomar~13. Their projected, intrinsic velocity dispersion of $\\sigma_\\mathrm{p} = 2.2\\pm 0.4$~km\\thinspace s$^{-1}$\\ implied a mass-to-light ratio $\\mathcal{M}/L_V= 40^{+24}_{-17}$, about one order of magnitude larger than the usual value for globular clusters. C\\^ot\\'e et al.\\ discussed at length all possible reasons for such an unusual result: (i) some velocity ``jitter'' among the red giants, (ii) a few binary stars, (iii) a non-standard mass function, (iv) process of dissolving into the Galactic halo through catastrophic tidal heating during a recent perigalacticon passage, or (v) the presence of a massive non-baryonic dark matter halo. It is worth emphasizing that, in C\\^ot\\'e et al., careful determination of the error bars made the usual mass-to-light ratio values for globular clusters $\\mathcal{M}/L_V\\sim 3$ at about two sigmas from the $\\mathcal{M}/L_V$ value obtained for Palomar~13. Because of this marginally significant and puzzling result from C\\^ot\\'e et al., some more spectroscopic data of stars in the field of Palomar~13 were acquired during the commissioning of the ESO/VLT instrument FLAMES. We present hereafter new high-quality and homogeneous radial velocities for 46 stars, 9 of them being members of Palomar~13, which provide new velocity dispersion and mass-to-light ratio values for this globular cluster. The remaining of this paper is as follows: Sect.~2 presents the observation and the data reduction, Sect.~3 discusses the membership of the stars, Sect.~4 gives the new velocity dispersion and corresponding mass-to-light ratio, and Sect.~5 discusses the plausible reasons for the difference between the present results and those obtained by C\\^ot\\'e et al. ", "conclusions": "" }, "0403/astro-ph0403666_arXiv.txt": { "abstract": "{ We present the COMBO-17 object catalogue of the Chandra Deep Field South for public use, covering a field which is $31\\farcm5 \\times 30\\arcmin$ in size. This catalogue lists astrometry, photometry in 17 passbands from 350 to 930~nm, and ground-based morphological data for 63,501 objects. The catalogue also contains multi-colour classification into the categories {\\it Star}, {\\it Galaxy} and {\\it Quasar} as well as photometric redshifts. We include restframe luminosities in Johnson, SDSS and Bessell passbands and estimated errors. The redshifts are most reliable at $R<24$, where the sample contains approximately 100 quasars, 1000 stars and 10000 galaxies. We use nearly 1000 spectroscopically identified objects in conjunction with detailed simulations to characterize the performance of COMBO-17. We show that the selection of {\\it quasars}, more generally type-1 AGN, is nearly complete and minimally contaminated at $z=[0.5,5]$ for luminosities above $M_B = -21.7$. Their photometric redshifts are accurate to roughly 5000~km/sec. Galaxy redshifts are accurate to 1\\% in $\\delta z/(1+z)$ at $R<21$. They degrade in quality for progressively fainter galaxies, reaching accuracies of 2\\% for galaxies with $R \\sim 22$ and of 10\\% for galaxies with $R>24$. The selection of stars is complete to $R \\sim 23$, and deeper for M stars. We also present an updated discussion of our classification technique with maps of survey completeness, and discuss possible failures of the statistical classification in the faint regime at $R\\ga 24$. ", "introduction": "The Chandra Deep Field South (CDFS) is one of the most well-studied patches of sky. It is the target of enormous observational efforts across a wide range of photon energies. The variety of imaging and spectroscopic data sets shall improve our understanding of fundamental processes in galaxy evolution. A large amount of public data are contributed by the {\\it Great Observatories Origins Deep Survey} (GOODS). This survey obtains deep images of the field using all of NASA's great space-based facilities: the Chandra X-ray observatory \\citep[CXO,][]{GiaCDFS}, the {\\it Advanced Camera for Surveys} (ACS) onboard the {\\it Hubble Space Telescope} \\citep[HST,][]{Giav04}, and the new infrared space telescope Spitzer. Further space-based observations include the {\\it Ultra Deep Field} (UDF) project targetting a small part of the field with a single ACS pointing, deep observations with ESA's X-ray observatory XMM-Newton (PI Bergeron), and the wider-area ACS imaging by the GEMS team \\citep{Rix04}. In this paper, we publish data and results from ground-based observations of the CDFS. Our project, COMBO-17, has targetted the CDFS among four other fields. They are all observed with the {\\it Wide Field Imager} \\citep[WFI,][]{WFI1,WFI2} at the MPG/ESO 2.2m-telescope on La Silla, Chile. This camera covers an area of more than $0.5\\degr \\times 0.5\\degr$, which is larger than the field initially observed from space by GOODS. The footprint of this larger WFI-based image is occasionally called {\\it Extended CDFS} or E-CDFS, but we just call it CDFS here. The purpose of the later GEMS images was to cover this larger area with HST resolution. COMBO-17 was mainly carried out to study the evolution of galaxies and their associated dark matter haloes at $z\\la 1$ as well as the evolution of quasars at $1\\la z\\la 5$. In order to obtain large samples of objects, four fields with a total area of $\\sim 1~\\sq\\degr$ were observed with a 17-band filter set covering the range of $\\lambda_{\\mathrm obs} \\approx 350 \\ldots 930$~nm. This provides {\\it very-low-resolution spectra} which allow a reliable classification into stars, galaxies and quasars as well as accurate photometric redshifts. This paper publishes the full COMBO-17 catalogue \\footnote{Catalogue and images are available at CDS and at the COMBO-17 website, http://www.mpia.de/COMBO/combo\\_index.html} on the CDFS with astrometry and 17-filter photometry \\footnote{In this paper, magnitudes are always used with reference to Vega as a zero point.} of 63,501 objects found on an area of $31\\farcm5 \\times 30\\arcmin$. We also include classification, photometric redshifts and restframe luminosities \\footnote{Throughout the paper we use $H_0 = h~\\times$ 100~km/(s~Mpc) in combination with $(\\Omega_m,\\Omega_\\Lambda)=(0.3,0.7)$ and $h=1$ for luminosity distances and restframe absolute magnitudes.} whereever the data permit their derivation. We believe, the classification is mostly reliable at $R\\la 24$, where the sample contains $\\sim 100$ QSOs, $\\sim 1000$ stars and $\\sim 10000$ galaxies. Wolf et al. (2001c) published an earlier version of the catalogue containing only astrometry and BVR photometry. The version published here contains the same set of objects with identical astrometry. However, after the photometry has been processed with our final procedures, we include all 17 passbands, classifications and redshifts. Our catalogue could be used directly to analyse aspects of galaxy evolution, and some results involving more COMBO-17 fields have already been published: Wolf et al. (2003a) studied the evolution of the galaxy luminosity function by spectral type from redshift 1.2 to 0.2. Bell et al. (2004) have focussed in particular on understanding the red sequence evolution over this redshift. Accurate photometric redshifts of QSOs allowed us to observe the evolution of faint AGN from redshift 5 to 1 (Wolf et al. 2003b) and calculate luminosity functions from the largest faint and unabsorbed AGN sample to date. Another obvious application is the selection of sub-samples for detailed observations, e.g. high-resolution spectroscopy, while relying on the knowledge of redshift and spectral type of targets. A first example drawn from this catalogue is the measurement of velocity dispersions for $z\\sim 1$ red sequence galaxies in the CDFS by van der Wel et al. (2004). A number of weak lensing studies took particular advantage of the accurate photometric redshifts provided by COMBO-17: Kleinheinrich et al. (2004) have used galaxy-galaxy lensing to study dark matter haloes of galaxies and their dependence on observed galaxy properties. Gray et al. (2004) have discussed the correlation of galaxy properties with the underlying dark matter density field, based on a weak lensing mass map obtained by Gray et al. (2002). Brown et al. (2003) derived the shear power sepctrum and constrained cosmological parameters from weak lensing and redshift distributions in COMBO-17. Heymans et al. (2004) have later removed intrinsic alignment signals based on our photometric redshifts. Bacon et al. (2004) have constrained the growth of dark matter density fluctuations with decreasing redshift. Taylor et al. (2004) have demonstrated the benefit of accurate redshifts through discovering a background galaxy cluster in projection behind a known cluster and estimating its mass purely from 3-D lensing analysis. Of course, the newly discovered cluster could also be confirmed independently from the redshift catalogue itself. A second purpose of this paper is to serve as a reference for the methodology of the classification and redshift estimation in COMBO-17. It is an update to the earlier and more detailed paper by Wolf, Meisenheimer \\& R\\\"oser (2001), hereafter WMR. In conjunction with WMR this paper provides a full description of the technique. We describe the performance of the classification and redshift estimation in the COMBO-17 data set as far as we can assess it at this time. We assume that our catalogue will only be useful if we provide estimates of completeness, contamination and accuracy of redshifts in the star, galaxy and quasar sample. We believe that our redshifts for galaxies are accurate within $\\sigma_z/(1+z) < 0.01$ at the bright end ($R<20$) where we also expect small outlier rates around 1\\%. Redshift errors increase towards fainter levels and exceed $\\sim$0.05 at $R>23.5$. Without NIR data faint galaxies at $z>1$ pose a tough challenge for our approach. We believe that the accuracy of our QSO redshifts is $\\sigma_z/(1+z)\\approx 0.015$ at all magnitudes where QSOs can be identified. In the future, we might be able to test the quality of the photometric redshifts more thoroughly. The ESO team of GOODS plans a large VIMOS programme to obtain low-resolution spectra of more than 5000 galaxies in the CDFS. Such a valuable resource would allow the most systematic test of photometric redshifts from deep images to date. A large number of redshifts have already been obtained by the VIRMOS VLT Deep Survey \\citep[VVDS,][]{LeF03} team in VIMOS GTO time. In this paper we briefly describe the COMBO-17 observations (Sect.~2) and data reduction (Sect.~3), followed by an update of our classification procedure and template choice over WMR (Sect.~4). In Sect.~5, we present the data structure of the catalogue for the CDFS, and Sect.~6 gives an overview of the object samples. In Sect.~7 we discuss completeness and redshift errors which we estimate from simulations of the survey. Finally, Sect.~8 discusses redshift errors in detail given the comparison with almost 1000 spectroscopic IDs of stars, galaxies and QSOs. Whenever we present numbers in this paper, we refer specifically to the CDFS dataset as it is published here, but whenever we talk about techniques in general, they are applied to all fields observed in COMBO-17. ", "conclusions": "Wolf, Meisenheimer \\& R\\\"oser (2001) showed that medium-band surveys deliver more accurate object classifications and photometric redshifts than broad-band surveys, while not consuming more telescope time. As a result, the COMBO-17 survey was started to obtain a large redshift catalogue of galaxies and AGN for evolutionary population studies, including weak lensing observations. In this paper, we have discussed in detail the quality of the classification and photometric redshifts of galaxies and quasars in COMBO-17. We have shown that the identification of stars is complete to $R\\la 23$ (deeper for M~stars). We have demonstrated that the identification of type-1 AGN is complete above luminosities of $M_B=-21.7$ at all redshifts from 0.5 to 5. The photometric redshifts of galaxies in COMBO-17 are better than 0.01 at bright magnitudes ($R<21$) and increase with photometric noise to $\\delta_z/(1+z) \\approx 0.06$ at $R=24$. Fainter than $R=24$, COMBO-17 is not particularly useful because the medium-bands are too shallow then. We have demonstrated that we routinely obtain photometric redshifts of quasars and luminous Seyfert-1 galaxies to an accuracy of $\\delta_z/(1+z) \\approx 0.015$. We have demonstrated that the medium-band approach indeed delivers the expected performance which motivated the survey COMBO-17. In this paper we now deliver the COMBO-17 database from one particular patch of sky to the community for public exploitation. The published database includes images and a catalogue with 63,501 objects. Classification and redshifts are typically reliable at $R<24$, where we find $\\sim 100$ quasars, $\\sim 1000$ stars and $\\sim 10000$ galaxies. We included the Chandra Deep Field South from the very beginning in the COMBO-17 project. A multitude of deep observations was expected across a wide range of photon energies, creating a unique data set for studies of galaxy evolution. The COMBO-17 approach has allowed us to get hold of a large redshift catalogue on the field. This includes the implicit selection of $\\sim 100$ luminous type-1 AGN with photometric redshifts as accurate as $\\sim 5000$~km/sec. We are now in a position to make this catalogue available for general use and hope to feed many more dedicated follow-up studies." }, "0403/astro-ph0403516_arXiv.txt": { "abstract": "Relativistic current sheets have been proposed as the sites of dissipation in pulsar winds, jets in active galaxies and other Poynting flux dominated flows. It is shown that the steady versions of these structures differ from their nonrelativistic counterparts because they do not permit transformation to a de~Hoffmann/Teller reference frame, in which the electric field vanishes. Instead, their generic form is that of a true neutral sheet: one in which the linking magnetic field component normal to the sheet is absent. Taken together with Alfv\\'en's limit on the total cross-field potential, this suggests plasma is ejected from the sheet in the cross-field direction rather than along it. The maximum energy to which such structures can accelerate particles is derived, and used to compute the maximum frequency of the subsequent synchrotron radiation. This can be substantially in excess of standard estimates. In the magnetically driven gamma-ray burst scenario, acceleration of electrons is possible to energies sufficient to enable photon-photon pair production after an inverse Compton scattering event. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403502_arXiv.txt": { "abstract": "The Six Degree Field Galaxy Survey (6dFGS) is a spectroscopic survey of the southern sky, which aims to provide positions and velocities of galaxies in the nearby Universe. When completed the survey will produce approximately 170000 redshifts and 15000 peculiar velocities. The survey is being carried out on the Anglo Australian Observatory's (AAO) UK Schmidt telescope, using the 6dF robotic fibre positioner and spectrograph system. We present here the adaptive tiling algorithm developed to place 6dFGS fields on the sky, and allocate targets to those fields. Optimal solutions to survey field placement are generally extremely difficult to find, especially in this era of large-scale galaxy surveys, as the space of available solutions is vast (2N dimensional) and false optimal solutions abound. The 6dFGS algorithm utilises the Metropolis (simulated annealing) method to overcome this problem. By design the algorithm gives uniform completeness independent of local density, so as to result in a highly complete and uniform observed sample. The adaptive tiling achieves a sampling rate of approximately 95\\%, a variation in the sampling uniformity of less than 5\\%, and an efficiency in terms of used fibres per field of greater than 90\\%. We have tested whether the tiling algorithm systematically biases the large-scale structure in the survey by studying the two-point correlation function of mock 6dF volumes. Our analysis shows that the constraints on fibre proximity with 6dF lead to under-estimating galaxy clustering on small scales ($<$1\\Mpc) by up to $\\sim$20\\%, but that the tiling introduces no significant sampling bias at larger scales. The algorithm should be generally applicable to virtually all tiling problems, and should reach whatever optimal solution is defined by the user's own merit function. ", "introduction": "\\label{tilingintro} \\begin{figure*} \\includegraphics[width=18cm]{Fig1.ps} \\caption{The 6dFGS targets show strong clustering on the sky, as can be seen in this equal--area (Aitoff projection) greyscale map of the surface density of targets. As the 6dF field covers an area of 25.5\\,deg$^2$ and has up to 150 fibres, an optimal surface density would be approximately 6 targets per deg$^2$. The large, and spatially complex, density variations about this optimum illustrate one of the major difficulties in tiling the 6dFGS.} \\label{greyscale} \\end{figure*} The advent of large-scale spectroscopic surveys, made possible by high multiplex spectroscopic systems, has necessitated the development of automated schemes for placing survey fields (`tiles') on the sky, and allocating survey targets to those fields. Adaptive tiling schemes take into account survey and instrument characteristics and provide efficient and optimal tile placement and target allocation. The recently completed 2dF Galaxy Redshift Survey (2dFGRS) successfully utilized adaptive tiling to obtain 221414 redshifts, using a 400 fibre spectrograph with a 2\\degree\\ field of view \\citep{Colless4}. The 2dFGRS covered 2000 deg$^2$ at a median depth of $\\bar{z}=0.11$. The Sloan Digital Sky Survey (SDSS) \\citep{York} aims to observe $\\sim$10$^6$ targets with a 640 fibre system and a 3\\degree\\ field of view, and is also employing adaptive tiling \\citep{Blanton2}. The SDSS will cover $\\sim$10000\\,deg$^2$ at a depth similar to the 2dFGRS. The 6dFGS is a redshift and peculiar velocity survey that will cover the 17000 deg$^2$ of the southern sky with $|b|>10$\\degree \\citep{Watson2,Saunders,Wakamatsu}. The survey is being carried out on the AAO's Schmidt telescope, using the 6dF automated fibre positioner and spectrograph system \\citep{Parker,Watson1}. 6dF can simultaneously observe up to 150 targets in a circular 5.7\\degree\\ field of view. Survey observations are made with two different gratings for each field. These two spectral ranges are spliced together as part of the redshifting process, resulting in single spectra that span the range from 3900\\AA\\ to 7500\\AA\\ , at a resolution of $R=1000$ at 5500\\AA\\ and a typical signal-to-noise ratio of $S/N\\sim10$. The goals of the survey are to map the positions and velocities of galaxies in the nearby Universe, providing new constraints on cosmological models, and a better understanding of the local populations of normal galaxies, radio galaxies, AGN and QSOs \\citep{Saunders}. The primary targets for the redshift survey are 113988 $K_s$-selected galaxies from the 2MASS near-infrared sky survey (\\citep{Jarrett}; {\\tt{http://www.ipac.caltech.edu/2mass/releases/allsky}}) down to $K_{tot}<12.75$ and with a median redshift $\\bar{z}=0.05$. The total magnitudes are estimated from the 2MASS isophotal $K_{20}$ magnitudes and surface brightness profile information \\citep{Jones}. Merged with the primary sample are 16 other smaller extragalactic samples, including targets selected from the HIPASS HI radio survey \\citep{Koribalski}, the ROSAT All Sky Survey of X-ray sources (\\cite{Voges1,Voges2}; {\\tt{http://heasarc.gsfc.nasa.gov/docs/rosat/ass.html}}), the IRAS Faint Source Catalogue (\\citep{Moshir}; {\\tt{http://irsa.ipac.caltech.edu/IRASdocs/iras.html}}), the DENIS near-infrared survey \\citep{Epchtein}, the SuperCosmos $b_J$ and $r_F$ optical catalogues \\citep{Miller}, the Hamburg-ESO QSO survey \\citep{Wisotzki} and the NVSS radio survey \\citep{Condon}. In total the survey will produce approximately 170000 redshifts. The 6dFGS peculiar velocity survey will consist of all early-type galaxies from the primary redshift survey sample that are sufficiently bright to yield precise velocity dispersions. These galaxies are observed at higher signal-to-noise ratio ($S/N>25$), in order to obtain velocity dispersions to an accuracy of 10\\%. Peculiar velocities will be obtained using the Fundamental Plane for early-type galaxies \\citep{Djorgovski,Dressler} by combining the velocity dispersions with the 2MASS photometry. Based on the high fraction of early-type galaxies in the $K_s$ sample and the $S/N$ obtained in our observations to date, we expect to measure distances and peculiar velocities for 10--15000 galaxies out to distances of at least \\mbox{$cz=15000$\\kms}. Observations have so far been made for $40\\%$ of the survey fields and completion is expected mid--2005. The data is non-proprietary and an Early Data Release for some 14000 objects can be accessed at {\\tt{http://www-wfau.roe.ac.uk/6dFGS/}}. This paper describes the adaptive tiling algorithm developed for the 6dFGS. It is organised in the following manner: \\S2 outlines the functional requirements for the tiling algorithm and the context in which it was developed; \\S3 gives a detailed explanation of the algorithm; \\S4 outlines the process of parameter selection and application of the algorithm to the 6dFGS catalogue; \\S5 presents an investigation of possible systematic effects introduced by the tiling, and their impact on subsequent analyses of survey data; \\S6 concludes with a summary of the tiling algorithm and its performance. \\begin{table} \\caption{The distribution of 6dFGS targets in terms of the numbers of neighboring targets within the fibre-button proximity exclusion limit. Only 60\\% of the catalogue are without close neighbours (as compared with $\\sim$90\\% in the SDSS), meaning a significant proportion have multiple close neighbours, the most extreme being one target with 40 neighbours within 5.71\\,arcmin.} \\begin{tabular}{crc} \\# Neighbours & \\# Targets & Sample fraction \\\\ 0 & 102252 & 59.2\\% \\\\ 1 & 43196 & 25.0\\% \\\\ 2 & 15695 & 9.1\\% \\\\ $\\ge3$ & 11604 & 6.7\\% \\\\ \\label{neighbours} \\end{tabular} \\end{table} ", "conclusions": "\\label{conc} Utilizing an optimization method based on simulated annealing, we have successfully developed an adaptive tiling algorithm to optimally place 6dFGS fields on the sky, and allocate targets to those fields. The algorithm involves a four-stage process: (i)~establishing individual target weights based on target surface density and sample observational priorities; (ii)~creating a database of all possible conflicts in allocating neighbouring targets closer than the radius of a 6dF fibre button; (iii)~creating an initial tiling by centering tiles on randomly selected targets, and then allocating targets to those tiles in order of decreasing numbers of neighbours and increasing separation from tile centres; (iv)~and finally, using the Metropolis method in randomly shifting the position of tiles, and then reallocating targets, to maximise the objective function of the tiling and hence provide an optimal tiling solution. In order to maximise the uniformity of sampling of the 6dFGS targets, we weight inversely with the surface density of 2MASS $K_s$ galaxies. Our results showed this gave us superior uniformity when compared with a simple uniform density weighting scheme, most noticeably in reducing the number of targets not allocated to tiles along the edges of the survey volume. Despite the challenges of highly clustered targets and large fibre buttons, tiling solutions generated using the algorithm are highly complete and uniform, and employ an efficient use of tiles. The tilings consistently give sampling rates of around 95\\%, with variations in the uniformity of sampling of less than 5\\%. Tiles typically have more than 90\\% of their available fibres allocated to targets. The algorithm has also proved itself highly flexible, able to perform on highly irregularly shaped distributions of targets. An analysis of the two-point correlation function, calculated from 6dF mock volumes tiled with the algorithm, revealed that the constraint on fibre proximity due to the large size of the fibre buttons produces a significant under-sampling of close pairs of galaxies on scales of 1 \\Mpc\\ and smaller; on larger scales, however, the tiling algorithm does not lead to any detectable sampling bias." }, "0403/astro-ph0403028_arXiv.txt": { "abstract": "{ In comparison with other well studied star formation regions, Taurus is unusual in several respects. (i) Its stellar initial mass function (IMF) peaks at relatively high mass ($\\sim 0.8 M_\\odot$), but contains very few stars much more massive than $1 M_\\odot$, and is relatively deficient in brown dwarfs. (ii) It has a higher binary fraction, particularly at large separations. (iii) Its core mass function is strongly peaked at a few $M_\\odot$, and the cores have extended envelopes and relatively low levels of turbulence. \\hspace{0.25cm} We present here the results of an ensemble of hydrodynamic simulations which suggest that the unusual stellar IMF in Taurus is a direct consequence of the unusual properties of the cores there. By following the collapse and fragmentation of cores having properties typical of Taurus, we find that roughly 50\\% of the objects formed in a core, predominantly the low-mass ones, are ejected from the core to form a population of low-mass stars and brown dwarfs with a flat mass function. The remaining objects form multiple systems within the core, accreting until their masses approach $1 M_\\odot$; this produces a population of intermediate-mass stars whose mass function peaks at $\\sim 0.8 M_\\odot$. Together these two populations reproduce the IMF in Taurus very well. This demonstrates, for the first time, a direct causal link between the core mass function and the stellar IMF in a star formation region. ", "introduction": "Stars form within dense molecular cores (e.g. Andr\\'{e} et al., 2000), and the densest and most centrally condensed cores, i.e. those closest to forming stars, are known as prestellar cores (Ward-Thompson et al., 1994). In Ophiuchus (Motte, Andr\\'e \\& Neri, 1998), Serpens (Testi \\& Sargent, 1998) and Orion (Motte et al., 2001), the mass function of pre-stellar cores is remarkably similar to the IMF for field stars and clustered star formation regions. This suggests that there is a simple mapping from the core mass function into the stellar IMF, with the mean masses of the stars forming within a core being proportional to the core mass. However, the details of how this works are still uncertain. In this paper we present numerical simulations of core collapse and fragmentation which demonstrate a causal link between the core mass function and the stellar IMF in the Taurus star formation region. We choose to study the Taurus region, because it is nearby and extended on the sky, and has therefore been studied in detail. Molecular-line mapping has yielded estimates of the Taurus core mass function (Onishi et al., 2002), and deep surveys of its stellar content have revealed the Taurus IMF down to the deuterium burning limit (Luhman et al., 2003a). ", "conclusions": "Taurus has an unusual stellar initial mass function and an unusual core mass function. We have modelled the hydrodynamical evolution of an ensemble of cores with masses based on the Taurus core mass function and levels of turbulence based on those observed in Taurus. We find that the unusual stellar IMF in Taurus can be explained as a direct result of the unusual core mass function and intrinsic core properties in Taurus. In each core an initially unstable, multiple system forms, with between 2 and 9 members. Typically, 2 or 3 objects are ejected before they can accrete a significant amount of material. These ejected low-mass stars and brown dwarfs constitute the flat, low-mass `tail' of the Taurus IMF. There are almost twice as many low-mass stars ($0.08 M_\\odot < M < 0.5 M_\\odot$) as brown dwarfs ($M < 0.08 M_\\odot$). The ejection velocities ($\\sim 1\\,-\\,2\\,{\\rm km}\\,{\\rm s}^{-1}$) are essentially independent of mass, so the low-mass stars are as dispersed as the brown dwarfs and there is no significant mass segregation. The remaining objects stay near the centre of the core and continue to accrete until their masses are $\\sim 0.8 M_{\\odot}$, by which stage there is not much mass left to accrete. These more massive stars constitute the Gaussian peak centred at $\\sim 0.8 M_\\odot$ in the Taurus IMF. Almost all of them are in binary or triple systems. Thus the simulated cluster of cores has an IMF very similar to that of Taurus, viz. a narrow Gaussian peak at $\\sim 0.8 M_\\odot$, and a flat tail at lower masses, extending into the brown dwarf regime. Further support for the hypothesis that the form of the IMF in Taurus is a direct result of the typical masses of the cores in Taurus comes from Goodwin et al. (in preparation) who find that a core with low levels of turbulence typically forms a number of objects approximately equal to the number of initial Jeans masses in the core (the initial Jeans mass being roughly $1 M_{\\odot}$). Thus lower-mass cores form fewer objects and are unable to populate the tail of the IMF as they do not eject significant numbers of brown dwarfs and low-mass stars. Higher-mass cores over-populate the tail through the ejection of more brown dwarfs and low-mass stars. In addition the stars that remain within a core have a larger resovoir of gas to accrete and become larger than the observed $0.8 M_{\\odot}$ peak. The agreement between the two histograms of Figure 3 (observation and theory) is remarkably close. Hence we believe we have shown, for the first time, a direct causal link between the core mass function of a star-forming region and the stellar IMF produced in that region." }, "0403/astro-ph0403672_arXiv.txt": { "abstract": " ", "introduction": "The aim of this article is to demonstrate the useful role that can be played by spectropolarimetric observations of young and evolved emission line stars that analyse the linearly polarized component in their spectra. At the time of writing, this demonstration has to be made on the basis of optical data since there is no common-user infrared facility, in operation, that offers the desired combination of spectral resolution and sensitivity. If the new ESO instrument, CRIRES, can be characterised to sufficient precision, it may become the first capable of performing such science. The case for high signal-to-noise (S/N) spectropolarimetry on the largest available telescopes has been made before at an earlier ESO conference (\\cite{Schmid}, \\cite{Donati}). Because the polarised fraction of the light received from astrophysical sources is often only of the order of 1\\% or so, it is clear that observations seeking to separate out and characterise this component suffer at least a 5-magnitude disadvantage relative to total light spectroscopy: expressed in S/N terms, one can only begin to achieve anything at S/N approaching 1~000. It was shown by both Schmid et al~\\cite{Schmid} and Donati et al~\\cite{Donati} that some of the most exciting results, particularly from the analysis of circular-polarised light, will only come as the achievable S/N rises to 100~000. Here we focus on what can be learned from linear spectropolarimetry alone at reasonably high spectral resolution (R of order 20000 is already valuable) and at $10^3 < $S/N$ < 10^4$. And we remind that the near infrared (1$\\umu$m--2$\\umu$m) has the potential to out-perform the optical as a domain to work in because of the greatly reduced interstellar obscuration at these wavelengths. ", "conclusions": "" }, "0403/astro-ph0403391_arXiv.txt": { "abstract": "We present the results from an automated search for damped \\lya\\ (DLA) systems in the quasar spectra of Data Release 1 from the Sloan Digital Sky Survey (SDSS-DR1). At $z\\approx 2.5$, this homogeneous dataset has greater statistical significance than the previous two decades of research. We derive a statistical sample of \\ndla\\ damped \\lya\\ systems ($>50$ previously unpublished) at $z>2.1$ and measure H\\,I column densities directly from the SDSS spectra. The number of DLA systems per unit redshift is consistent with previous measurements and we expect our survey has $>95\\%$ completeness. We examine the cosmological baryonic mass density of neutral gas $\\Omega_g$ inferred from the damped \\lya\\ systems from the SDSS-DR1 survey and a combined sample drawn from the literature. Contrary to previous results, the $\\Omega_g$ values do not require a significant correction from Lyman limit systems at any redshift. We also find that the $\\Omega_g$ values for the SDSS-DR1 sample do not decline at high redshift and the combined sample shows a (statistically insignificant) decrease only at $z>4$. Future data releases from SDSS will provide the definitive survey of DLA systems at $z\\approx 2.5$ and will significantly reduce the uncertainty in $\\Omega_g$ at higher redshift. ", "introduction": "It has now been two decades since the inception of surveys for high redshift galaxies through the signature of damped \\lya\\ (DLA) absorption in the spectra of background quasars \\citep{wolfe86}. Owing to large neutral hydrogen column densities $\\N{HI}$, these absorption lines exhibit large rest equivalent widths ($W_\\lambda > 10$\\AA) and show the Lorentzian wings characteristic of quantum mechanic line-damping. Through dedicated surveys of high and low redshift quasars with optical and ultraviolet telescopes, over 300 damped \\lya\\ systems have been identified. These galaxies span redshifts $z=0$ (the Milky Way, LMC, SMC) to $z=5.5$ where the opacity of the \\lya\\ forest precludes detection \\citep{songaila02}. Statistics of the DLA systems impact a wide range of topics in modern cosmology, galaxy formation, and physics. These include studies on the chemical enrichment of the universe in neutral gas \\citep{pettini94,pro03b}, nucleosynthetic processes \\citep{lu96,phw03}, galactic velocity fields \\citep{pw97}, the molecular and dust content of young galaxies \\citep{vladilo98,ledoux03}, star formation rates \\citep{wpg03}, and even constraints on temporal evolution of the fine-structure constant \\citep{webb01}. Perhaps the most fundamental measurement from DLA surveys, however, is the evolution of the cosmological baryonic mass density in neutral gas \\ohi\\ \\citep[][ hereafter PMSI03]{storrie00,rao00,peroux03}. Because the DLA systems dominate the mass density of neutral gas from $z=0$ to at least $z=3.5$, a census of these absorption systems determines directly $\\Omega_g$. These measurements express global evolution in the gas which feeds star formation \\citep{pei95,mathlin01} and are an important constraint for models of hierarchical galaxy formation \\citep[e.g.][]{spf01,nagamine04a}. The most recent compilation of damped \\lya\\ systems surveyed in a `blind', statistical manner combines the effects of observing programs using over 10 telescopes, 10 unique instruments, and the data reduction and analysis of $\\approx 10$ different observers (PMSI03). In short, the results are derived from a heterogeneous sample of quasar spectra derived from heterogeneous quasar surveys. While considerable care has been paid to collate these studies into an unbiased analysis, it is difficult to assess the completeness and potential selection biases of the current sample. These issues are particularly important when one aims to address the impact of effects like dust obscuration \\citep{ostriker84,fall93,ellison01}. In this paper we present the first results in a large survey for damped \\lya\\ systems drawn from a homogeneous dataset of high $z$ quasars with well-defined selection criteria. Specifically, we survey the quasar spectra from Data Release 1 of the Sloan Digital Sky Survey (SDSS-DR1) restricting our search to SDSS-DR1 quasars with Petrosian magnitude $r' < 19.5$\\,mag. The DR1 sample alone (the first of five data releases from SDSS) offers a survey comparable to -- although not strictly independent from -- the efforts of 20 years of work. We introduce algorithms to automatically identify DLA candidates in the fluxed (i.e.\\ non-normalized) quasar spectra and perform Voigt profile analyses to confirm and analyze the DLA sample. This survey was motivated by a search for `metal-strong' DLA systems like the $z$=2.626 damped \\lya\\ system toward FJ$0812+32$ \\citep{phw03}. A discussion of the `metal-strong' survey will be presented in a future paper (Herbert-Fort et al.\\ 2004, in preparation). This paper is organized as follows. In $\\S$~2, we present the quasar sample and discuss the automatic DLA candidate detection. In $\\S$~3, we present the Voigt profile fits to the full sample. We present a statistical analysis in $\\S$~4 and a summary and concluding remarks are given in $\\S$~5. \\clearpage ", "conclusions": "In this paper, we have introduced an automated approach for identifying DLA systems in the SDSS quasar database. We have applied our method to the Data Relase 1 quasar sample and have identified a statistical sample of \\ndla\\ DLA systems including $>50$ previously unpublished cases. Remarkably, the SDSS Data Release 1 exceeds the statistical significance of the previous two decades of DLA research at $z \\approx 2.5$. More importantly, this sample was drawn from a well defined, homogeneous dataset of quasar spectroscopy. We present measurements of the number per unit redshift $n(z)$ of the DLA population and the contribution of these systems to the cosmological baryonic mass density in neutral gas $\\Omega_g$. Although the SDSS-DR1 sample does not offer a definitive assessment of either of these quantities, future SDSS data releases will provide a major advancement over all previous work. Our measurements of $n(z)$ are consistent with previous results suggesting a high completeness level for our DLA survey of the SDSS-DR1. We find $\\Omega_g$ increases with redshift to at least $z=3$ and is consistent with increasing to $z=4$ and beyond. This latter claim, however, is subject to significant uncertainty relating to sample size. Perhaps the most important result of our analysis is that the full DLA sample no longer shows significantly fewer DLA systems with large $\\N{HI}$ at $z>3.5$. This contradicts the principal result of PMSI03 from their analysis of the pre-SDSS DLA compilation. Apparently, their maximum likelihood approach failed to adequately assess uncertainty related to sample size. With the inclusion of only 6 new DLA, we no longer find that Lyman limit systems with $\\N{HI} < 2 \\sci{20} \\cm{-2}$ are required in an analysis of $\\Omega_g$. Before concluding, we offer several additional criticisms of the PMSI03 analysis and the role of sub-DLA systems. First, these authors assumed a three parameter $\\Gamma$-function for the column density frequency distribution of absorption systems with $\\N{HI} > 10^{17.2} \\cm{-2}$, \\\\ $f(N) = (f_*/N_*) (N/N_*)^{-\\beta} {\\rm e}^{-N/N_*}$. Although this function gives a reasonable fit to the column density frequency distribution of the DLA systems, it is not physically motivated\\footnote{In fact this curve does not smoothly connect to the power-law derived for quasar absorption lines with $\\N{HI} < 10^{17.2} \\cm{-2}$.} and, more importantly, places much greater emphasis on sub-DLA than other functions (e.g.\\ a broken power-law). Future assessments must include other functional forms to examine this systematic uncertainty. Second, the authors did not fit for the normalization of the distribution function $f_*$. The uncertainty in this parameter could easily contribute an additional $>50\\%$ to the error budget. Third, their treatment did not account for sample variance; the uncertainties these authors reported were severe underestimates. Finally (and perhaps most importantly), a recent analysis of a sub-DLA sample by \\cite{dessauges03} has shown that these absorption systems have very high ionization fractions (see also Howk \\& Wolfe 2004 in preparation). Although these absorption systems may ultimately make an important contribution to the total H\\,I mass density of the universe, they are intrinsically different from the DLA systems. Indeed, a more appropriate title for this sub-set of Lyman limit systems is the `super-LLS'. This gas -- in its present form -- cannot contribute to star formation and is unlikely to be directly associated with galactic disks or the inner regions of protogalactic `clumps'. Any interpretation of results related to the super-LLS must carefully consider these points \\citep[e.g.][]{maller03,peroux03}." }, "0403/astro-ph0403358_arXiv.txt": { "abstract": "Explaining the origin of the orbit of 2000~CR$_{105}$ ($a\\sim 230$~AU, $q\\sim 45$~AU) is a major test for our understanding of the primordial evolution of the outer Solar System. Gladman et al$.$~(2001) showed that this objects could not have been a normal member of the scattered disk that had its perihelion distance increased by chaotic diffusion. In this paper we explore four seemingly promising mechanisms for explaining the origin of the orbit of this peculiar object: (i) the passage of Neptune through a high-eccentricity phase, (ii) the past existence of massive planetary embryos in the Kuiper belt or the scattered disk, (iii) the presence of a massive trans-Neptunian disk at early epochs which exerted tides on scattered disk objects, and (iv) encounters with other stars. Of all these mechanisms, the only one giving satisfactory results is the passage of a star. Indeed, our simulations show that the passage of a solar mass star at about 800~AU only perturbs objects with semi-major axes larger than roughly 200~AU to large perihelion distances. This is in good agreement with the fact that 2000~CR$_{105}$ has a semi-major axis of 230~AU and no other bodies with similar perihelion distances but smaller semi-major axes have yet been discovered. The discovery of 2003 VB$_{12}$, ($a=450$~AU, $q=75$~AU) announced a few days before the submission of this paper, strengthen our conclusions. ", "introduction": "\\label{intro} The trans-Neptunian population of small bodies is usually divided in two categories, the Kuiper belt and the scattered disk, although the partition between the two is not precisely defined. In Morbidelli et al. (2003) we have introduced a partitioning based on the dynamics of orbits in the current Solar System. We called {\\it scattered disk} the region of the orbital space that can be visited by bodies that have encountered Neptune within a Hill's radius at least once during the age of the Solar System, assuming no substantial modification of the planetary orbits. We then called {\\it Kuiper belt} the complement of the scattered disk in the $a>30$~AU region. The bodies that belong to the scattered disk in this classification scheme do not provide us with any significant clue about the primordial architecture of the Solar System. This is because their current orbits can be achieved by purely dynamical evolution in the current planetary system from objects that started in nearly-circular nearly-coplanar orbits in Neptune's zone. The opposite is true for the orbits of the Kuiper belt objects. All bodies in the Solar System must have been formed on orbits typical of an accretion disk (e.g. with very small eccentricities and inclinations). Therefore, the fact that most Kuiper belt objects have a non-negligible eccentricity and/or inclination reveals that some excitation mechanism, which is no longer at work, was active in the past. In this respect, particularly interesting are the Kuiper belt bodies with large semi-major axis ($a>50$~AU), such as 2001~QW$_{297}$ ($a=51.3$~AU, $q=39.5$AU, $i=17.1^\\circ$), 2000~YW$_{134}$ ($a=58.4$~AU, $q=41.2$AU, $i=19.8^\\circ$), 1995~TL$_8$ ($a=52.5$~AU, $q=40.2$AU, $i=0.2^\\circ$), 2000~CR$_{105}$ ($a=230$~AU, $q=44.4$~AU, $i=22.7^\\circ$) and, last discovered, 2003~VB$_{12}$ ($a=531$~AU, $q=74.4$~AU, $i=11.9^\\circ$; Brown et al$.$~2004). We call these objects {\\it extended scattered disk} objects for three reasons: {\\it (i)} they do not seem to belong (some caution is needed because of the uncertainties in the orbital elements of these objects) to the scattered disk according to our definition but are very close to its boundary (Gladman et al., 2001; Emel'yanenko et al., 2003; Morbidelli et al., 2004); {\\it (ii)} some of these bodies have sizes of several hundred kilometers, suggesting that they formed much closer to the Sun, where the accretion timescale was sufficiently short (Stern, 1996) and were subsequently transported to these current locations; {\\it (iii)} the lack of objects with $q>41$~AU and $5050$~AU region. However, in all cases, the semi-major axis was smaller than 200~AU. Therefore, Gomes's mechanism implies the existence of bodies with $q\\!sim\\!45\\,$AU spread from $a\\!\\sim\\!50\\,$AU to $\\sim\\!200\\,$AU, but no objects has been discovered at the small semi-major axis end of this range. This, in spite of the fact that observational biases favor the discovery of small semi-major axis objects. Indeed, 2000~CR$_{105}$ is special for a couple of reasons. Until the recent discovery of 2003~VB$_{12}$, it had the largest semi-major axis of the extended scattered disk, by a large margin. It also had a significantly larger perihelion distance than any other extended scattered disk object. Although it is possible that 2000~CR$_{105}$ is just an outlaying member of the extended scattered disk, the fact that no objects with perihelion distance comparable to that of 2000~CR$_{105}$ but with a smaller $a$ had been discovered seemed significant to us. This is particularly true considering that observational biases sharply favors the discovery of objects with smaller semi-major axes. Thus, we were motivated to look for dynamical mechanisms that preferentially raised the perihelion distance of scattered disk objects at large semi-major axis. The discovery of 2003~VB$_{12}$ came a few days before the submission of this paper, and confirmed that our investigation was well motivated. In fact, the orbit of this body definitely falls beyond the distribution produced in Gomes model. Some of the mechanisms investigated in this paper have been already suggested by Gladman et al. (2001), but never have been quantitatively explored. In Section~\\ref{nep_ecc} we consider the case where Neptune was more eccentric in the past, as proposed by Thommes et al$.$~(1999). It is obvious that a more eccentric Neptune would produce a extended scattered disk, but it is not known, {\\it a priori}, what eccentricity would be required to produce objects on 2000~CR$_{105}$-like orbits, and over what timescale. In Section~\\ref{planet} we investigate the effects of the presence of terrestrial mass planet(s) in the Kuiper belt or in the scattered disk, as proposed by Morbidelli and Valsecchi~(1997) and Brunini and Melita~(2002). In Section~\\ref{dtides} we propose a new model for the origin of 2000~CR$_{105}$, in which the tides raised by a massive disk beyond $\\sim 70$~AU increased the perihelion distance of high inclined scattered disk objects. Finally, in Section~\\ref{senc} we investigate the stellar passage scenario. This scenario has been first proposed by Ida et al. (2000) to explain the structure of the inner Kuiper belt. Although we disagree that the all of the sculpting of the Kuiper belt could be due this mechanism (see Levison et al$.$, 2004), it is still possible that a more gentle encounter could have formed objects like 2000~CR$_{105}$. ", "conclusions": "We have analyzed with numerical simulations four seemingly promising mechanisms for explaining the origin of the peculiar extended scattered disk object 2000~CR$_{105}$: (i) a high eccentricity phase of Neptune, (ii) the existence of a rogue planet in the Kuiper belt or in the scattered disk, (iii) the tide exerted by a massive and dynamically cold trans-Neptunian disk, and (iv) the passage of a star near the Solar System. Of these, only the early passage of a Solar-mass star at about 800~AU from the Sun appears satisfactory. This is also the only scenario that we have studied that can easily explain the origin of the newly found object 2003~VB$_{12}$. Another scenario for the origin of extended scattered disk objects has been proposed by Gomes (2003a, 2003b). In it, a small fraction of the objects in an early massive scattered disk population permanently acquire a large perihelion distance during the outer migration of Neptune. This mechanism predicts that there should be large-$q$ objects with semi-major axis all over the range from $\\sim\\!50$ to $\\sim\\!200\\,$AU. The fact that the extended scattered disk bodies with the largest perihelion distances, 2000~CR$_{105}$ and 2003~VB$_{12}$, both have $a>200$~AU, argues against a scenario like that of Gomes. Since observational biases (given an object's perihelion distance and absolute magnitude, and a survey's limiting magnitude of detection) sharply favors the discovery of objects with small semi-major axis, we believe that it would be unlikely that the first two discovered body with $q>44$~AU had $a\\!>\\!200$~AU, if the real semi-major axis distribution in the extended scattered disk were skewed toward small $a$. An yet, Gomes's mechanism and all of the mechanisms we have studied, except for the passing star, lead to such a distribution. Indeed, it is intuitive that creating a distribution where the objects with the largest $a$ also have the largest $q$ requires a perturbation `from the outside', whose magnitude decreases with decreasing heliocentric distance. Thus, there are not many alternatives to the stellar encounter scenario. In the current galactic environment, the closest stellar encounter that should occur over the age of the Solar System is at $\\sim 900$~AU (Garcia-Sanchez et al., 2001). This should typically happen with a star that is about 1/10th the mass of the Sun. Therefore, the encounter with a Solar-mass star at $\\sim 800$~AU at early times requires that the environment in which that Sun formed was significantly denser, such as that of a stellar cluster (Bate et al., 2003). In the future the existence of 2000~CR$_{105}$, 2003~VB$_{12}$, and their cohorts my supply important clues to the exact environment in which the Sun and Solar system formed. On a final note, since its discovery, passing stars have been used to capture objects into the Oort cloud (Oort 1950). Indeed, the process invoked for the Oort cloud is identical to the one employed here. Thus, if a passing star is indeed responsible for the formation of objects like 2000~CR$_{105}$ and 2003~VB$_{12}$, then it is perhaps more appropriate to characterize these objects as the inner edge of the Oort cloud rather than the outer edge of the scattered disk. Indeed, recent simulations of the Oort cloud in a star cluster (Eggers~1997; 1998; Fernandez and Brunini~2000) actually produce objects like 2003~VB$_{12}$." }, "0403/astro-ph0403444_arXiv.txt": { "abstract": "{Combining precise $B$,$V$ photometry and radial velocities, we have been able to derive a firm orbital solution and accurate physical parameters for the newly discovered eclipsing binary HD~23642 in the Pleiades open cluster. The resulting distance to the binary and therefore to the cluster is 132$\\pm$2~pc. This closely confirms the distance modulus obtained by classical main sequence fitting methods ($m - M$=5.60 or 132 pc), moving cluster techniques and the astrometric orbit of Atlas. This is the first time the distance to a member of the Pleiades is derived by orbital solution of a double-lined eclipsing binary, and it is intended to contribute to the ongoing discussion about the discordant Hipparcos distance to the cluster. ", "introduction": "The distance to the Pleiades has been derived by main-sequence fitting methods several times over the years, and consistently found to cluster around $m - M$=5.60~$\\pm$0.04 (e.g. Turner 1979, Meynet et al. 1993), corresponding to a distance of 132~($\\pm$2) pc. Values of $m - M$=5.52 (127~pc) by Mitchell \\& Johnson (1957) and $m - M$=5.75 (141~pc) by Eggen (1950) bracket the range of published distance moduli. Therefore, it came as a surprise the shorter distance of 116~$\\pm$3.3~pc ($m - M$=5.33~$\\pm$0.06) derived from the parallaxes of 54 Pleiades members when the results of the Hipparcos astrometric mission became available (van Leeuwen \\& Hansen Ruiz 1997). The $\\sim$0.3~mag difference has far reaching consequences in many areas of astrophysics, and it soon prompted extensive observational and theoretical work to account for it. \\begin{figure*}[!t] \\centerline{\\psfig{file=Gb231_f1.ps,width=16.7 cm,angle=270}} \\caption[]{Our photo-electric $V$ (top) and $B$ (bottom) observations of HD~23642 are phase plotted in the upper panel following the ephemeris of Table~2. The lower panel displays our radial velocities from Table~1 (filled circles) and those from Pearce (1957, squares) and Abt (1958, crosses). The curves over plot the orbital solution given in Table~2.} \\end{figure*} An anomalous abundance of Helium was discussed as a possible explanation of the large difference between Hipparcos and ground-based photometric distances to the Pleiades (Mermilliod et al. 1997a). The required helium over-abundance ($Y$=0.37) is however too large to be a feasible explanation according to Pinsonneault et al. (1998). A sub-solar metallicity could reconcile the distances, but the metallicity of the Pleiades has been measured several times and with different methods, which generally converged to a mean value of [Fe/H]=0.00$\\pm$0.03 (e.g. Stauffer et al. 2003). Castellani et al. (2002) were able to obtain a good fit of the Pleiades main sequence with their theoretical isochrones (with updated physical inputs) and the Hipparcos distance, employing a sub-solar metallicity of [Fe/H]=--0.15. Working differentially with a sample of nearby G and K stars, Percival et al. (2003) presented a re-calibration of photometric metallicity indexes and argued that they support a sub-solar metallicity of the Pleiades in spite of the spectroscopic evidences (e.g. Boesgaard \\& Friel 1990) for a solar one, therefore removing the difference between main-sequence fitting and Hipparcos distances. However, Stauffer et al. (2003) remarked about the anomalous blue colors of K stars in the Pleiades, that they ascribed to fast rotation and spotted surfaces as a consequence of the young age of the cluster. Narayanan \\& Gould (1999) derived a distance modulus to the Pleiades of $m - M$=5.58~($\\pm$0.18) using a variant of the moving cluster method, the gradient in the radial velocity of the cluster members in the direction of the proper motion of the cluster. Their techniques relies on the assumption that the velocity structure of the Pleiades is not significantly affected by rotation. Narayanan \\& Gould concluded that the errors in the Hipparcos parallaxes toward Pleiades are spatially correlated over angular scales of a few degrees, with an amplitude up to 2~mas. Almost simultaneously, however, van Leeuwen (1999) compared the results on 9 clusters and concluded instead that the Hipparcos parallaxes of the Pleiades were basically unaffected by systematics and that the distance modulus to the latter is $m - M$=5.37~($\\pm$0.07). Makarov (2002) suggested that, while Hipparcos parallaxes are overall of excellent quality, those for the Pleiades suffered from data analysis procedures that could have introduced systematics when a rich and bright cluster was crossing one of the two Hipparcos fields of view while the other was essentially deprived of stars, as in the case of Pleiades. From intermediate Hipparcos data and under some assumptions, Makarov has recomputed the distance to the Pleiades as 129~$\\pm$3~pc. F. van Leeuwen (private communication) is currently working on a further iteration on original Hipparcos reductions that should assess and quantify is any systematics affect the Pleiades data. The question about the distance to the Pleiades has not yet been firmly solved, and to tackle it new, robust and independent approaches (as much geometric as possible) are required. A first one, advocated by Pacz\\'{y}nski (2003), combines astrometric and spectroscopic observations of Atlas (HD~23850), one of the brightest members of the Pleiades, which is an astrometric binary with a period of 291~day, a semi-major axis of 12.9~mas and 0.246 eccentricity. Pan et al. (2004) did not have the radial velocities, but they anyway derived the distance by combining the astrometric orbit with a mass-luminosity relation. They derived a distance of 135$\\pm$2~pc, which is in close agreement with the results of main-sequence fitting methods. Another method involves double-lined eclipsing binaries (SB2 EB), which are the distance indicators now providing the most reliable distances to Magellanic Clouds and other galaxies in the Local Group. The recent discovery by Torres (2003) that HD~23642 in the Pleiades is an SB2 eclipsing binary (the only one so far known in the cluster) prompted us to use it to measure the distance to the Pleiades. In this {\\em Letter} we present accurate new $B$ and $V$ photoelectric photometry and radial velocities of HD~23642, and use them to derive an orbital solution and infer the distance to the star, and therefore to the cluster. ", "conclusions": "" }, "0403/hep-th0403073_arXiv.txt": { "abstract": "Dirac-Born-Infeld type effective actions reproduce many aspects of string theory classical tachyon dynamics of unstable Dp-branes. The inhomogeneous tachyon field rolling from the top of its potential forms topological defects of lower codimensions. In between them, as we show, the tachyon energy density fragments into a p-dimensional web-like high density network evolving with time. We present an analytic asymptotic series solution of the non-linear equations for the inhomogeneous tachyon and its stress energy. The generic solution for a tachyon field with a runaway potential in arbitrary dimensions is described by the free streaming of noninteracting massive particles whose initial velocities are defined by the gradients of the initial tachyon profile. Thus, relativistic particle mechanics is a dual picture of the tachyon field effective action. Implications of this picture for inflationary models with a decaying tachyon field are discussed. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403608_arXiv.txt": { "abstract": "At $z=0.1055$, the gamma-ray burst GRB 031203 is one of the two nearest GRBs known. Using observations from the Very Large Array (VLA) and \\chandralong\\, we derive sub-arcsecond localizations of the radio and X-ray afterglow of this GRB. We present near-infrared observations of the supernova SN 2003lw, which exploded in the host galaxy of the GRB 031203. Our deep, high resolution Magellan/PANIC data establish that this SN is spatially coincident with the radio and X-ray localizations of the afterglow of GRB 031203 to sub-arcsecond precision, and is thus firmly associated with the GRB. We use image differencing to subtract the bright emission from the host galaxy, and measure the time evolution of the SN between $\\sim5$ and $\\sim50$ days after the GRB. The resulting light curve has a shape which is quite different from that of the two SNe previously associated with GRBs, SN 1998bw and SN 2003dh. With SN 2003lw securely associated with this burst, we confirm that all three GRBs with redshifts $z<0.3$ were accompanied by SN explosions. ", "introduction": "The emerging association between long-duration gamma-ray bursts (GRBs) and type Ic supernovae (SNe Ic) is perhaps the most significant breakthrough in our understanding of GRBs, and may also provide new insights into the physics of core-collapse SNe and the deaths of massive stars. The initial strong evidence for this connection came from the spatial and temporal coincidence between GRB 980425 and SN 1998bw (Galama et al. 1998). More recently, this picture was convincingly affirmed by the detection of SN features in the optical afterglow spectrum of GRB 030329 (Stanek et al. 2003; Hjorth et al. 2003; Matheson et al. 2003). Both of these milestone discoveries resulted from the study of the nearest GRBs yet identified, GRB 980425 at $z=0.0085$ and GRB 030329 at $z=0.1685$. Indeed, each relatively rare occurrence of a GRB at low redshift ($z<0.3$) provides a unique opportunity for further study of the GRB-SN connection. GRB 031203\\footnotemark~was detected by IBIS on board the INTEGRAL spacecraft on 2003 December 3, at 22:01:28 UT (Gotz et al. 2003). A fading X-ray afterglow was discovered by XMM-Newton on 2003 December 4 (Santos-Lleo et al. 2003; Campana et al. 2003; Rodriguez-Pascual et al. 2003) and found to be consistent with the location of a radio transient (Frail 2003; Soderberg, Kulkarni, \\& Frail 2003). Prochaska et al. (2003a; 2003b; 2004) identified the host galaxy of the X-ray and radio transients, determined its redshift ($z=0.1055$) and studied its properties in detail. Recently, several groups (Bersier et al. 2004; Tagliaferri et al. 2004a; Thomsen et al. 2004; Cobb et al. 2004) reported optical photometric and spectroscopic observations of this GRB, which apparently reveal the signatures of an associated SN (designated SN 2003lw; Tagliaferri et al. 2004b). In spite of the low redshift of this event, its low Galactic latitude (less than $5^{\\circ}$ from the plane) and the resulting Galactic extinction ($E[B-V]=1.04$; Schlegel, Finkbeiner, \\& Davis 1998), as well as its bright host galaxy, make the study of this event challenging. \\footnotetext{The possible classification of this event as an X-ray flash (XRF) was debated in recent literature (e.g., Prochaska et al. 2004; Watson et al. 2004; Thomsen et al. 2004). However, the high-energy observations presented by Sazonov, Lutovinov, \\& Sunyaev (2004) conclusively show that this event does not fit any of the commonly used definitions of XRFs.} To overcome these difficulties, we have undertaken near infrared (NIR) observations. The results of this effort are reported in this {\\it Letter}. Coordinated radio and X-ray observations, which enabled us to probe the total energy output of this sub-energetic GRB, are reported elsewhere (Soderberg et al. 2004). ", "conclusions": "In this {\\it Letter}, we have presented $J$-band observations of SN 2003lw. The sensitivity and resolution of our data enabled us to pinpoint the location of SN 2003lw within its host galaxy, and to show it is consistent with sub-arcsecond localizations of the radio and X-ray afterglow of GRB 031203, thus confirming the association of these two events. The precise NIR localization of this event also puts it within $0.2$~kpc from the host galaxy center. The $J$-band light curve of SN 2003lw shows a rapid initial rise ($5-7$ days after the GRB) and evidence for bright emission more than $50$ days after the GRB. The fast early rise of SN 2003dh, associated with GRB 030329, has been interpreted by Woosley \\& Heger (2003) and Mazzali et al. (2003) as evidence for asymmetry in the explosion. A thorough investigation of this possibility will probably require an analysis of our data in combination with other extensive data sets of optical and NIR photometry and spectroscopy collected by other groups (e.g., Thomsen et al. 2004; Tagliaferri et al. 2004; Bersier et al. 2004; Cobb et al. 2004). Cobb et al (2004) have recently reported $I$ and $J$-band observations of this event, obtained with the SMARTS 1.3m telescope. A direct comparison between our observations and the SMARTS data, obtained on numerous epochs, is complicated by the fact that these authors do not present the light curve of SN 2003lw. Instead, they plot the temporal evolution of the combined light of the SN and its bright host galaxy, derived from aperture photometry, which shows considerable scatter. It is thus hard to say whether their data show the same early fast rise we detect. Our observation that SN 2003lw had similar flux levels 7 and 50 days after the GRB is consistent with the reported SMARTS data. Furthermore, comparing their $I$-band data with model light curves of SN 1998bw, these authors arrive at the conclusion that the light curve shape of SN 2003lw does not resemble that of SN 1998bw. This is in accord with our analysis of the $J$-band data (Cobb et al. 2004 do not attempt to compare their $J$-band data with a model of SN 1998bw). It appears that both our observations and the Cobb et al. (2004) data set suggest that SN 2003lw had a light curve quite unlike that of SN 1998bw; with a fast rise to maximum, which appears broader, and perhaps showing a secondary peak in the IR. The above discussion provides further evidence for the diversity of SNe associated with GRBs [see, e.g., Thomsen et al. (2004) and Lipkin et al. (2004) for recent reviews]. With SN 2003lw, all three nearby GRBs are firmly associated with SNe, apparently supporting suggestions made, e.g., by Podsiadlowski et al (2004), that all long GRBs are accompanied by SNe. It may well be the case that the focus of future studies should now move from proving the association between SNe and GRBs to an attempt to characterize the properties of this population of SNe. Such studies may provide new clues about the progenitors and engines of GRBs, by requiring viable GRB models to be able to produce the large quantities of nickel derived from the SN observations, as well as valuable insights into possible explosion mechanisms of core-collapse SNe, which may involve GRB-like aspherical effects [e.g., Khokhlov et al. (1999), see Gal-Yam et al. (2004) for further discussion]. Low-redshift GRBs, expected to be localized in larger numbers by the upcoming {\\it SWIFT} mission, as well as systematic studies of local core-collapse SNe (e.g., Berger et al. 2003; Soderberg et al. 2004, in preparation; Stockdale et al. 2004) will probably shed more light on these intriguing questions." }, "0403/astro-ph0403114_arXiv.txt": { "abstract": "We report on observations of the blazar W Comae (ON+231) with the Solar Tower Atmospheric Cherenkov Effect Experiment (STACEE), a wavefront-sampling atmospheric Cherenkov telescope, in the spring of 2003. In a data set comprising 10.5 hours of ON-source observing time, we detect no significant emission from W Comae. We discuss the implications of our results in the context of the composition of the relativistic jet in W Comae, examining both leptonic and hadronic models for the jet. We derive 95\\% confidence level upper limits on the flux at the level of 1.5--3.5$\\times 10^{-10}$ cm$^{-2}$ s$^{-1}$ above 100 GeV for the leptonic models, or 0.5--1.1$\\times 10^{-10}$ cm$^{-2}$ s$^{-1}$ above 150 GeV for the hadronic models. ", "introduction": "The current catalog of extragalactic TeV gamma-ray sources consists of blazars, which are among the brightest and most rapidly variable objects in the sky. The unusual observational properties of blazars are usually explained in terms of relativistic bulk motion of the emitting region in a jet parallel to the line of sight \\citep{up95}. The continuum emission of blazars is typically nonthermal and contains two broad peaks, one at lower energies (radio to X-ray) and one at higher energies (keV to TeV). Although the low-energy peak is generally assumed to result from synchrotron radiation from high-energy electrons in the jet, the mechanism for the production of the high-energy radiation is still a subject of debate, and several competing models exist. In ``leptonic'' jet models, the high-energy radiation is produced by inverse Compton scattering of low-energy photons from a population of electrons and/or positrons; these models are favored for the well-studied blazars Mrk 421 and 501 \\citep{ca99,konopelko03}. In the synchrotron self-Compton model (SSC), a single population of electrons and positrons produce the target photon field via synchrotron radiation, and subsequently upscatter the synchrotron photons to gamma-ray energies \\citep{bm96}. There may also be an external Compton (EC) component of target photons from the accretion disk or ambient medium \\citep{dsm92,sbr94}. Alternatively, in ``hadronic'' jet models, protons play a central role. In hadronic models the high-energy radiation is attributed to photomeson processes \\citep{mannheim93} or synchrotron radiation from protons or muons, as in the synchrotron proton blazar (SPB) model \\citep{mp00,aharonian00,mucke03}. The study of hadronic models was originally motivated by the hypothesis that the highest-energy cosmic rays are produced in blazar jets \\citep{mannheim95}. The BL Lac object W Comae (ON+231) may provide an excellent test case for hadronic jet models \\citep{bmr02}. The transition between the low-energy and high-energy peaks in the continuum of W Comae appears clearly in X-ray data taken by the BeppoSAX satellite \\citep{tagliaferri00}. These high-quality observations of the transition region place tight constraints on leptonic models, requiring the predicted gamma-ray emission to cut off sharply above 100 GeV. In contrast, hadronic models may allow for significant emission above 100 GeV. Observations by the EGRET detector aboard the Compton Gamma Ray Observatory show a hard power law spectrum (photon spectral index $\\alpha = 1.73 \\pm 0.18$) extending up to about 10 GeV with no sign of any cutoff \\citep{hartman99}. Yet the object has not been detected at energies above 300 GeV, despite repeated observation by the Whipple 10-m instrument \\citep{horan03}. At a redshift of $0.1$, absorption of gamma rays in this energy range by pair production \\( \\gamma \\gamma \\longrightarrow e^+e^- \\) against the extragalactic background light (EBL) may be significant, but only at energies above about 500 GeV \\citep{primack99,primack01,ms01,aharonian01}, so the intrinsic emission spectrum of W Comae should be directly observable at energies lower than this. The Solar Tower Atmospheric Cherenkov Effect Experiment (STACEE) currently operates above an energy threshold of about 100 GeV for gamma rays. W Comae has been observed in the past with previous versions of the STACEE detector \\citep{theoret-thesis}. New observations of W Comae were carried out in the spring of 2003 by the most recent version of the detector, STACEE-64. ", "conclusions": "The STACEE-64 observations in the spring of 2003 were made at a lower energy threshold than any other atmospheric Cherenkov telescope has yet attained for W Comae. STACEE detects no significant emission from W Comae, resulting in 95\\% CL upper limits on the integral flux above this threshold in various hadronic emission models at the level of $10^{-10}$ cm$^{-2}$ s$^{-1}$. While leptonic models predict a flux which falls below this level, extrapolations of the best-fit EGRET power law, and some synchrotron-proton hadronic models, predict an integral gamma-ray flux above the energy threshold close to, or exceeding, the upper limit from STACEE observations. Additional STACEE observations planned for the spring of 2004 ought either to exclude these models at a significantly higher confidence level, or to detect gamma-ray emission from W Comae if these models provide an adequate description of the source." }, "0403/astro-ph0403264_arXiv.txt": { "abstract": "Using the SPectrometer for Infrared Faint Field Imaging (SPIFFI) on the ESO VLT, we have obtained $J$, $H$, and $K$ band integral field spectroscopy of the $z = 2.565$ luminous submillimeter galaxy SMM\\,J14011+0252. A global spectrum reveals the brighter of this spatially resolved system's two components as an intense starburst that is remarkably old, massive, and metal-rich for the early epoch at which it is observed. We see a strong Balmer break implying a $\\geq 100\\,{\\rm Myr}$ timescale for continuous star formation, as well as nebular emission line ratios implying a supersolar oxygen abundance on large spatial scales. Overall, the system is rapidly converting a large baryonic mass into stars over the course of only a few hundred Myr. Our study thus adds new arguments to the growing evidence that submillimeter galaxies are more massive than Lyman break galaxies, and more numerous at high redshift than predicted by current semi-analytic models of galaxy evolution. ", "introduction": "Deep imaging in the optical and near-IR has made it possible to constrain both the evolution of the cosmic star formation rate density (e.g., Madau et al. 1996) and its time integral, the growth of the cosmic stellar mass density (e.g., Dickinson et al. 2003). Short-wavelength studies give an incomplete picture of these trends, however, since an important population of high-redshift galaxies is too obscured to be easily detected. These submillimeter galaxies (SMGs; see Blain et al. 2002 and references therein) have large IR luminosities that are predominantly generated by star formation rather than accretion \\citep{barg01,alma03}; as a result, they contribute substantially to cosmic star formation \\citep{barg00}. Moreover, as the strikingly different appearances of the Hubble Deep Field at $0.83\\,{\\rm \\mu m}$ \\citep{will96} and $850\\,{\\rm \\mu m}$ \\citep{hugh98} exemplify, SMGs are rarer and forming stars much more intensely than typical optically selected systems. To understand the evolutionary state of SMGs, we have conducted a detailed rest-frame optical case study using SPIFFI, the SPectrometer for Infrared Faint Field Imaging \\citep{tecz00,eise00,eise03}. Our target is SMM\\,J14011+0252 (hereafter J14011), a $z = 2.565$ galaxy lying behind the $z = 0.25$ cluster A1835 and the second SMG to have an optical counterpart identification \\citep{barg99} validated by CO interferometry \\citep{fray99}. Recent maps of J14011's molecular gas have led to divergent conclusions about its intrinsic size: \\citet{ivis01} suggest that a large background source undergoes a factor ${\\cal M} \\sim 2.5$ magnification by A1835 as a whole, while \\citet{down03} argue that a fortuitously located cluster member further boosts magnification of a small background source to ${\\cal M} \\sim 25$. We defer a full discussion of lensing to a future paper; here, we focus primarily on the lensing-independent conclusions of high age and metallicity that can be drawn from J14011's global spectrum alone. Throughout the paper we assume a flat $\\Omega_\\Lambda = 0.7$ cosmology with $H_0 = 70\\,{\\rm km\\,s^{-1}\\,Mpc^{-1}}$. ", "conclusions": "" }, "0403/astro-ph0403578_arXiv.txt": { "abstract": "\\noindent Obtaining simultaneous radio and X--ray data during the outburst decay of soft X--ray transients is a potentially important tool to study the disc -- jet connection. Here we report results of the analysis of (nearly) simultaneous radio (VLA or WSRT) and {\\it Chandra} X--ray observations of XTE~J1908+094 during the last part of the decay of the source after an outburst. The limit on the index of a radio -- X--ray correlation we find is consistent with the value of $\\sim0.7$ which was found for other black hole candidates in the low/hard state. Interestingly, the limit we find seems more consistent with a value of 1.4 which was recently shown to be typical for radiatively efficient accretion flow models. We further show that when the correlation--index is the same for two sources one can use the differences in normalisation in the radio -- X--ray flux correlation to estimate the distance towards the sources if the distance of one of them is accurately known (assuming black hole spin and mass and jet Lorentz factor differences are unimportant or minimal). Finally, we observed a strong increase in the rate of decay of the X--ray flux. Between March 23, 2003 and April 19, 2003 the X--ray flux decayed with a factor $\\sim$5 whereas between April 19, 2003 and May 13, 2003, the X--ray flux decreased by a factor $\\sim$750. The source (0.5--10 keV) luminosity at the last {\\it Chandra} observation was ${\\rm L\\approx 3\\times10^{32} (\\frac{d}{8.5 kpc})^2 erg\\,s^{-1}}$. ", "introduction": "\\label{intro} Low--mass X--ray binaries (LMXBs) are binary systems in which a $\\approxlt 1\\,M_{\\odot}$ star transfers matter to a neutron star or a black hole. These systems form one of our main windows on the physical processes taking place around black holes and hence they can provide us with information about the fundamental properties of spacetime. One reason for this is that the great majority of Galactic black hole candidates (BHCs) are found in transient LMXB systems. Over the last few years it has become apparent that jets are an integral and energetically important part of these BHC systems (especially) when these systems are in the so called low/hard state (\\pcite{2001MNRAS.322...31F}; \\pcite{2001MNRAS.327.1273S}). Recently, it was found that there exists a correlation between the radio and X--ray flux in the low/hard state of several BHCs over 3--4 decades in X--ray flux showing that there must be some form of disc--jet coupling (\\pcite{2003A&A...400.1007C}; \\pcite{2003MNRAS.344...60G}). \\scite{2003MNRAS.343L..59H}, \\scite{falcke2003}, and \\scite{merheinzdim2003} review the disc--jet connection in terms of different accretion disc and jet models. \\scite{merheinzdim2003}, building on previous work of \\scite{2003MNRAS.343L..59H}, showed that inefficient accretion flow models can reproduce the observed radio -- X--ray correlation index for the initial parameter space they covered. \\scite{2003A&A...397..645M} showed that the jet--model explaining the observed X--rays in terms of synchrotron emission from the jet of \\scite{2001A&A...372L..25M} can reproduce both the observed correlation index as well as the normalisation. \\scite{2003MNRAS.343L..99F} used the observed radio -- X--ray correlation for BHCs to argue that there is no need to advect energy across a black hole event horizon in order to explain the observed difference in quiescent luminosity between the neutron star and BHC transient systems as was proposed by e.g.~\\scite{2001ApJ...553L..47G}. An important assumption in the work of \\scite{2003MNRAS.343L..99F} is that the observed radio -- X--ray correlation holds down to X--ray luminosities as low as L$_X \\sim10^{30-32}$ erg s$^{-1}$. XTE~J1908+094 was discovered serendipitously during {\\it RXTE} observations of the soft gamma--ray repeater SGR~1900+14 by \\scite{2002IAUC.7856....1W}. The source flux is absorbed (${\\rm N_H\\sim2.3\\times10^{22} cm^{-2}}$), the spectrum is well--fit with a hard power--law with a photon index of 1.55. Subsequent {\\it BeppoSAX} observations (\\pcite{2002IAUC.7873....1I}; \\pcite{2002A&A...394..553I}) confirmed both the hard spectrum (the source was detected up to 250 keV) and the high Galactic absorption. A broadened iron emission line was present in spectra extracted from both the RXTE and the BeppoSAX observations. In 't Zand et al. (2002) presented strong evidence for a low/hard -- high/soft state change. The fact that the source displayed both a low/hard and a high/soft state during the outburst is confirmed by the timing and spectral analysis of the RXTE/PCA observations by \\scite{2002xrb..confE..11G} (see also \\pcite{gogus2004}). A radio counterpart was discovered by \\scite{2002IAUC.7874....1R} whereas a near--infrared counterpart was found by \\scite{2002MNRAS.337L..23C}. These authors also found that the optical upper limits (\\pcite{2002ATel...86....1W} and \\pcite{2002IAUC.7877....4G}) are fully consistent with the near--infrared colours of and the high extinction towards the source. In this paper we report the findings of our (nearly) simultaneous Very Large Array (VLA)\\footnote{The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.} and Westerbork Synthesis Radio Telescopes (WSRT)\\footnote{The Westerbork Synthesis Radio Telescope is operated by the ASTRON (Netherlands Foundation for Research in Astronomy) with support from the Netherlands Foundation for Scientific Research NWO} radio and {\\it Chandra} X--ray observations of XTE~J1908+094 during the last part of the decay of the source after an outburst. ", "conclusions": "We have obtained (nearly) simultaneous VLA and WSRT radio and {\\it Chandra} X--ray observations of the BHC XTE~J1908+094 during the decay after an outburst. We find that: \\newline {\\it (i)} Limits on a power--law correlation between radio and X--ray flux suggest that the power--law index may be larger than that found before in other BHCs.\\newline {\\it (ii)} The rate of decay increases from a factor of $\\sim$5 in $\\sim$25 days to a factor of $\\sim$750 in $\\sim$25 days. \\newline {\\it (iii)} The source spectrum hardens during the decay. \\newline Below we will discuss these findings in more detail. Previously, \\scite{2003A&A...400.1007C} and \\scite{2003MNRAS.344...60G} found that there is a correlation between the radio and X--ray flux over more than 4 orders of magnitude. The index of the power--law relation they fitted was consistent with being the same for several sources at a value of 0.7$\\pm$0.1 (Gallo et al.~2003). If we assume that the radio flux of the source on April 19, 2003 was at the 1~$\\sigma$ level of the WSRT upper limit the index was 1.5$^{+0.45}_{-0.3}$ for a two--point power--law decay for XTE~J1908+094. This is consistent at the $\\sim3\\sigma$ level with the value of 0.7 found before. However, there are some caveats. This index was determined over a limited range in X--ray flux and using very few measurements only. Furthermore, we assumed that the radio spectral index is flat with $\\alpha=0$ (S$_\\nu\\propto\\nu^{\\alpha}$). Finally, the radio flux on Qpril 19, 2003 may have been much lower than what we have assumed; this would make the correlation index steeper. Previously, \\scite{2003MNRAS.342L..67M} found an index of $\\sim$1.4 for the radio -- X--ray correlation over a small range in X--ray flux in the neutron star system 4U~1728--34. Possibly, XTE~J1908+094 is a neutron star as well. However, \\scite{2002A&A...394..553I} and \\scite{2002xrb..confE..11G} argue strongly in favour of a BHC nature for XTE~J1908+094 on the basis of the observed outburst X--ray spectral and timing properties (see also Gogos et al.~2004, submitted), but obviously a dynamical mass estimate showing the mass of the compact object to be more than 3 M$_\\odot$ would settle the issue. If we follow the line of reasoning laid--out in \\scite{2003MNRAS.343L..99F} but taking ${\\rm L_{Radio}\\propto L_X^{1.4}}$ instead of ${\\rm L_{Radio}\\propto L_X^{0.7}}$ we find that ${\\rm L_{Jet}\\propto L_X}$. Hence, the ratio between jet and accretion power (which is assumed to be tracked by the X--ray luminosity) remains the same as the source flux decays. We note that the fact that the index was found to be 1.4 for a neutron star system does not affect the conclusion of \\scite{2003MNRAS.343L..99F} that the difference in quiescent X--ray luminosity between BHCs and neutron star soft X--ray transients can be explained without the need of advection of energy across the event horizon as long as the index for the BHCs is 0.7. Recently, \\scite{2003MNRAS.343L..59H} showed that the index of 0.7 follows naturally for several jet--models if one assumes that those jet models are scale invariant (i.e.~the Schwarzschild radius, r$_s$, is the only relevant length scale for jet formation). Building on the work of \\scite{2003MNRAS.343L..59H}, \\scite{merheinzdim2003} also showed that other indexes for the radio -- X--ray correlations could be found, i.e.~for both a gas and a radiation pressure dominated disc the index would be $\\sim$1.4, whereas it would be close to 0.7 for radiatively inefficient accretion flows. \\scite{2003A&A...397..645M} show that the model explaining part of the X--ray emission as jet synchrotron emission (\\pcite{2001A&A...372L..25M}) can explain the 0.7 index as well as the normalisation of the radio -- X--ray correlation. Hence, it seems that in XTE~J1908+094, during the part of the decay that we covered with our radio and X--ray observations, a standard geometrically thin optically thick disc plus a corona could have been present. Why accretion would proceed via a geometrically thin disc in XTE~J1908+094 whereas in other sources it is thought that the standard disc is not present in the low/hard state is unclear. Perhaps it has something to do with the luminosity levels at which the various sources are observed so far, but since the distance to XTE~J1908+094 is ill--constrained (\\pcite{2002A&A...394..553I} argue that the distance must be larger than 3 kpc) the source luminosity is not well known. From Fig.~\\ref{xdecay} it is clear that the rate of decay increased enormously after April 12, 2003. Such a steep decrease has been observed before for several BHCs and neutron star soft X--ray transients (e.g.~\\pcite{1997ApJ...491..312C}). In neutron star systems this has been interpreted as evidence for the onset of the propeller effect (\\pcite{1998ApJ...499L..65C}), however, since such a drop in luminosity seems to be common for BHCs as well, this interpretation may need to be revised (\\pcite{2003MNRAS.341..823J}). Finally, we find that the X--ray spectrum hardens between the first and second X--ray observation. Spectral hardening is often observed during/just after a transition to the low/hard state (cf.~\\pcite{2001ApJ...563..229T}). However, the findings of Gogos et al.~(2004) show that XTE~J1908+094 was already in the low/hard state several months before the first {\\it Chandra} observation was made. Perhaps the source changed from the low/hard state back to a soft state in between the {\\it RXTE} and the {\\it Chandra} observations. Such a change would be consistent with the fact that the limit on the radio -- X--ray correlation index is close to 1.4. We conclude that more observations at these low flux levels (and likely low luminosities) are necessary to determine the behaviour of these sources when they return back to quiescence." }, "0403/astro-ph0403052_arXiv.txt": { "abstract": "The High Energy Stereoscopic System (\\hess) - is a system of four, 107~m$^{2}$ mirror area, imaging Cherenkov telescopes under construction in the Khomas Highland of Namibia (1800~m asl). The \\hess~ system is characterised by a low threshold ($\\sim$ 100~GeV) and a $\\sim$1\\% Crab flux sensitivity resulting from the good angular resolution and background rejection provided by the stereoscopic technique. The first two telescopes are operational and first results are reported here. The remaining two telescopes (of \\hess~Phase-I) will be commissioned early in 2004. ", "introduction": "\\begin{figure} \\begin{center} \\includegraphics[height=19pc]{sensitivity.eps} \\end{center} \\caption{Expected point-source sensitivity of \\hess~Phase I compared to that of the VERITAS array\\cite{VERITAS} (reproduced from \\cite{HESS_ICRC01}).} \\label{fig_mc} \\end{figure} \\hess~is an array of 4 identical telescopes arranged in a square with 120~m sides. Each telescope has a focal length of 15 m and a 13~m diameter. The reflectors comprise 380 quartz-coated round facets (60~cm diameter), arranged with Davies-Cotton optics. The \\hess~cameras consist of 960 pixels of $0.16^\\circ$ angular size providing a total field of view of $5^\\circ$. A pixel consists of a photomultiplier tube (PMT) with a Winston cone light collector. The PMTs (Photonis XP2960) are organised into {\\em drawers} of 16 PMTs with associated read-out electronics. All triggering and read-out electronics are contained inside the camera body. Monte-Carlo simulations predict a sensitivity for the 4 telescope system of around 1\\% of the Crab flux (5$\\sigma$, 50 hours) and an energy threshold of $\\approx$100 GeV (see Fig.~\\ref{fig_mc}). ", "conclusions": "Two of the four telescopes of \\hess~Phase-I are complete and operational. The optical and mechanical performance of these telescopes meets all specifications. $\\gamma$-ray observations of the Crab confirm the simulated performance of a single \\hess~telescope. We confirm the BL-Lac PKS 2155-304 as a strong VHE $\\gamma$-ray source. The full \\hess~system should begin operations early in 2004." }, "0403/astro-ph0403347_arXiv.txt": { "abstract": "s{We discuss some of the most unusual active galactic nuclei (AGN) discovered to date by the Sloan Digital Sky Survey (SDSS): the first broad absorption line quasar seen to exhibit He\\,{\\sc ii} absorption, several quasars with extremely strong, narrow UV Fe\\,{\\sc ii} emission, and an AGN with an unexplained and very strange continuum shape. } ", "introduction": "The Sloan Digital Sky Survey (York et al. 2000; Fukugita et al. 1996; Gunn et al. 1998; Hogg et al. 2001; Stoughton et al. 2002; Smith et al. 2002; Pier et al. 2003) is obtaining optical spectra for $\\sim$10$^5$ quasars over $\\sim$$\\frac{1}{4}$ of the entire sky. Through careful target selection (Richards et al. 2002) and sheer size, the SDSS includes numerous AGN with unconventional properties. The high-quality, moderate-resolution SDSS spectra can be used to set the stage for the detailed multiwavelength studies often needed to understand interesting quasar subclasses. We illustrate this fact using several unusual AGN included in the SDSS Second Data Release (Abazajian et al. 2004). ", "conclusions": "" }, "0403/astro-ph0403171_arXiv.txt": { "abstract": "s{ We show that within a recently developed nonlocal, chiral quark model the critical densities for a phase transition to color superconducting quark matter under neutron star conditions can be low enough that these phases occur in compact star configurations with masses below $1.4~M_\\odot$. We study the cooling of these objects in isolation for different values of the gravitational mass and thus different composition and structure of the interior. Our equation of state allows for a 2SC phase with a large quark gap $\\Delta \\sim 100~$MeV for $u$ and $d$ quarks of two colors, a normal quark matter phase and their coexistence in a mixed phase within the hybrid star interior. We argue that, if the phases with unpaired quarks were allowed, the corresponding hybrid stars would cool too fast to describe the neutron star cooling data existing by today. We incorporate other attractive channels permitting a weak pairing of the residual quarks which remained unpaired in the 2SC phase and demonstrate that the model does not contradict the cooling data if the weak pairing gaps are of the order of $0.1~$ MeV. } ", "introduction": "In the recent paper \\cite{bgv2004}, hereafter BGV, we have reinvestigated the cooling of neutron stars (NS) within a purely hadron model, i.e., ignoring the possibility of quark cores in NS interiors. We have demonstrated that the NS cooling data available by today can be well explained within the {\\em \"Nuclear medium cooling scenario\"}, i.e., if one includes medium effects in the emissivity and takes into account a suppression of the $3P_2$ neutron gap. In a subsequent work \\cite{bgv2004q} we have shown that this result does not exclude the possibility that neutron stars might possess large quark matter cores that extend up to more than half of the star radius. Such a hybrid structure gives room for a whole variety of additional scenarios of compact star cooling which fall into two classes: either nuclear and quark matter phases have similar cooling behavior (homgeneous cooling) or the faster cooling of the one phase is compensated by the slower cooling of the other (inhomogeneous cooling). In the present contribution we will report on our results within the former, homgeneous cooling scenario of hybrid stars and what implications the comparison with present-day cooling data may provide for the EoS and transport properties of quark matter. ", "conclusions": "" }, "0403/astro-ph0403492_arXiv.txt": { "abstract": "\\noindent If the causality condition [the speed of sound always remains less than that of light in vacuum, i. e., $v \\leq c = 1$] is imposed on the spheres of homogeneous energy density, the `ratio of the specific heats', $\\gamma \\leq 2.59457$, constraints the compaction parameter, $u [\\equiv (M/a)$, mass to size ratio in geometrized units] of the dynamically stable configurations $ \\leq 0.34056 $ [corresponding to a surface redshift $(z_a ) \\leq 0.771$]. Apparently, The maximum value of $u$ obtained in this manner belongs to an absolute upper bound, and gives: (i) The maximum value for static neutron star masses as $5.4 M_\\odot$, if we substitute the density at the surface of the configuration equal to the average nuclear density, $E = 2 \\times 10^{14}$ g\\, cm$^{-3}$ [e.g. {\\it Nature}, {\\bf 259}, 377 (1976)]. (ii) However, if the density of the static configuration is constrained to the value $1.072 \\times 10^{14}$ g\\, cm$^{-3}$, by imposing the empirical result that the minimum rotation period of the fastest rotating pulsar known to date, PSR 1937 + 21, is 1.558 ms, the maximum mass value for static neutron stars exceed upto $7.4 M_\\odot$. These masses have important implications for the massive compact objects like Cyg X-1, Cyg XR-1, and LMC-X3 etc., which may not, necessarily, represent black holes. (iii) The minimum rotation periods for a static $1.442 M_\\odot$ neutron star to be 0.3041 ms. (iv) A suitable stable model of ultra-compact objects [$u \\ > (1/3)$] which has important astrophysical significance. ", "introduction": "% Incompressible fluid spheres of uniform energy density $E$ in General Relativity were first discussed by Schwarzschild.$^1$ The importance of this solution in General Relativistic stellar structures is apparent, because it gives an absolute upper limit on compaction parameter, $u (\\equiv M/a$, mass to size ratio of entire configuration in geometrized units) $ \\leq (4/9) $ for any regular static solution in hydrostatic equilibrium.$^2$ Chandrasekhar$^{3,4}$ discussed the condition of hydrostatic equilibrium under the small adiabatic perturbations, and showed that for each value of the compaction parameter, corresponding to the {\\em compressible} homogeneous spheres, there exists a critical (minimum) value of the ``ratio of specific heats'', $\\gamma (= \\gamma_{crit}$) such that for $\\gamma \\ < \\gamma_{crit}$, the configuration becomes dynamically unstable. For the limiting case of the compaction parameter approaching the Schwarzschild limit ($u = 4/9), \\gamma$ becomes infinity. For dynamically stable superdense objects like neutron stars one may expect a finite value of $\\gamma$. However, for such objects the equations of state are not well known [empirically] beyond the density $ \\cong 10^{14} $ g\\, cm$^{-3},^5$ and one can only extrapolate the equations of state (available in the literature)$^6$ far beyond this density. As a way out, one can impose some restrictions upon the known physical quantities, such that, the speed of sound inside the configuration, \\begin{displaymath} v \\equiv \\sqrt {(\\partial {P} / \\partial {E})_s} \\end{displaymath} (Where $P$ is the pressure, $E$ is the energy density and $s$ stands for specific entropy) does not exceed the speed of light in vacuum, i.e., $v \\leq c = 1$ (in geometrized units), and obtain an upper bound on stable neutron star masses.$^{7-9}$ In the present paper, we have obtained an upper bound on compaction parameter ($u \\leq 0.34056$ corresponding to a surface redshift of 0.771) for the compressible homogeneous spheres,$^{3,4}$ by imposing constraint on the ``ratio of specific heats'', $\\gamma [ \\leq 2.59457)]$, compatible with causality ($v \\leq 1$) and dynamical stability. This value of the compaction parameter is an absolute maximum because, for an assigned value of $\\gamma$, the maximum compactness would correspond to a compressible uniform density sphere, and can be used to obtain an upper bound on neutron star masses, as well as the minimum rotation period of a $1.442 M_\\odot$ neutron star (the maximum mass of the neutron star accurately known at present).$^{10}$ ", "conclusions": "An absolute upper bound on compaction parameter, $u \\leq 0.34056 $ [or the surface redshift $ \\leq 0.771$], compatible with causality and the ratio of the specific heats, $\\gamma \\leq 2.59457$, is obtained by using the dynamically stable compressible homogeneous sphere. This upper limit of compaction parameter gives (i) The maximum static mass of conventional model of neutron stars [taking $E = 2 \\times 10^{14}$ g\\, cm$^{-3}]^{7,8}$ as $5.4 M_\\odot$. This is greater than $4.8 M_\\odot$ considered as an upper limit earlier. (ii) The maximum mass of static neutron star exceeds to the value of $7.387 M_\\odot$ which is greater than the upper limit of $5.3 M_\\odot$ for neutron stars (so called $Q$-star models) obtained by Hochron, Lynn and Selipesky.$^{26}$ This may have important implications for the heavy compact objects like Cyg X-1, Cyg XR-1, and LMC-X3 which may not, necessarily, be black holes. (iii) The minimum rotation periods for a static $1.442 M_\\odot$ neutron star to be 0.3041 ms with a uniform energy density $E$ as $2.813 \\times 10^{15}$ g\\, cm$^{-3}$. (iv) A causally consistent and dynamically stable model of ultra-compact objects [$u \\ > (1/3)$] which has important astrophysical significance [see, e.g. Ref. 16 and references therein]." }, "0403/astro-ph0403201_arXiv.txt": { "abstract": "Protostellar sources in star forming regions are responsible for driving jets with flow velocities ranging between 300 and 400 km s$^{-1}$. This class of jets consists of highly collimated outflows which include thermal knots with number densities estimated to be greater than that of their ambient medium. On the other hand, extragalactic FR I jets consist of light fluid with low Mach number burrowing through a denser medium as the magnetized jets radiate nonthermal emission. Both protostellar and extragalactic jets are believed to be launched by accretion disks. Here we consider a jet model in which the characteristics common to both protostellar and extragalactic jets are used to explain the origin of nonthermal filaments in the Galactic center region. We argue that these filaments are analogous to FR I extragalactic sources but are launched by embedded young stars or clusters of stars in star-forming regions. ", "introduction": "It has been 20 years since the discovery of the nonthermal radio filaments (NRFs) associated with the Galactic center Arc was first reported (Yusef-Zadeh, Morris, \\& Chance 1984). These observations showed evidence of linear, magnetized features running perpendicular to the Galactic plane. A number of NRFs with similar characteristics to the prototype NRFs have been discovered in the intervening years (Yusef-Zadeh 2003 and references therein). Several models have suggested that the filaments trace the illuminated component of a large-scale poloidal magnetic field pervasive throughout the Galactic center region. However, the presence of a number of filaments oriented at large angles to the normal to the Galactic plane does not support the above interpretation and indicates a different origin. Unlike most models that predict a global, static geometry of the magnetic field around the Galactic center, the model described here argues for a local origin. In the proposed picture, the NRFs originate in star-forming regions. The filaments behave like jets extracting mass and energy from embedded young stars or clusters of stars as the jets propagate in a dense ISM of the Galactic center region. A more detailed account of this model will be given elsewhere. ", "conclusions": "" }, "0403/astro-ph0403037_arXiv.txt": { "abstract": "The principal results of daily observations with the RATAN-600 radio telescope of X-ray binary with relativistic jets microquasar SS433 in 1986--2003 are presented. We have measured the flux densities at 0.96, 2.3, 3.9, 7.7, 11.2 and 21.7 GHz in different sets, duration from a week to some months. In general there are 940 observations of SS433 and more than 4500 flux density measurements in the period. Observations show that radio spectra are well fitting by a power law. The mean spectral index remained the same, $-0.60\\pm0.14$ during almost 20 years at least, and mean accuracy of the index determination was better than 0.1 in our multi-frequency observations, i.e. it was higher than in the intensive two-frequency monitoring of SS433 with the three-element GBI interferometer. Flux density data and spectra `on-line' plotting are accessible on the CATS data base site: http://cats.sao.ru/. \\keywords {X-rays: binaries -- stars: flare -- stars: individual: SS433 -- jets -- radio continuum: stars -- monitoring } ", "introduction": "The X-ray binaries (XBs) have long been studied in radio band. After an identification of Cyg X-3 in the beginning of the 70-s it became evident that XBs could have powerful variable radio emission. A sample of X-ray binaries with relativistic jets that Mirabel {\\&} Rodriguez (1999) named microquasars consists of 15--20 objects. The brightest of them were observed actively under a monitoring program of XBs with the RATAN-600 radio telescope (Trushkin 2000, Trushkin \\& Bursov 2001). SS433 --- a bright variable emission star --- was identified by Clark and Murdin (1978) with a rather bright compact radio source 1909+048 located in the center of a supernova remnant W50. When in 1979 mobile optical emission lines were discovered in the spectrum of this bright star SS433 --- a radio source 1909+04 (Margon et al. 1979), it became apparent that a new class of objects in the Galaxy was found. At the same time Spencer (1979) was the first to discover an extended structure: a compact core and 1 arcsec long aligned jets in the radio image of SS433. At present such a structure in microquasars is commonly named a radio jet. Different data do indicate a presence of a very narrow (about 1$\\degr$) collimated beam at least in X-ray and optical ranges. At present there is no doubt that SS433 is related to W50. SS433 is probably a stellar remnant of a SN that exploded in a binary system about 30000~years ago. An explosion of one of the components did not destroy the binary. A distance to SS433 of 4.8 kpc was later determined by different ways including the direct measurement of proper motions of the jet radio components. As is now known, this unique variable X-ray, IR and radio source is an eclipsing binary cosisting of a compact object and an early type massive star. The system has a pair of opposite relativistic (in which matter is ejected at 0.26c) jets, which show themselves in different ways in different ranges from X-ray to radio. The VLBI (Very Long Baseline Radio Interferometry) observations of the jets proper motion showed that they consist of separate blobs that move ballistically at the same velocity 0.26c. Resultant from the 164-day precession of a thick accreting disk the jets rotate along conic surface and look like structures similar to a twin corkscrew on VLA radio maps (Hjellming {\\&} Johnston 1981). There are strong evidences for believing that this binary system consists of a relativistic star with an accreting disk and a massive optical component filling its Roche lobe. The supercritical accretion onto the relativistic star is likely to lead to initiation of the SS433 main feature --- two opposite jets of matter moving from the accretion disk poles at a velocity of about a quarter of the speed of light. The recent intensive spectral investigation made it possible to determine rather precisely the SS433 mass function and to estimate the relativistic component mass. It is equal to 11~$\\pm $~5~M$_{\\sun}$ (Gies et al. 2002a,b) what suggests that this is a black hole. Absorption lines in the SS433 spectrum belonging to the optical component are almost identical with those in the spectrum of an evolved star of a mid-A spectral class with a mass of 19~$\\pm $~7~M$_{\\sun}$. These jets that had been first discovered in spectral optical observations are seen also in X-ray and radio wavelengths. Activity phenomena in SS433 give rise to strong variability of all its electromagnetic spectrum. As is seen from the jet X-ray images and from an oblong form of the radio remnant W50, the jet structure extends up to size scales of $1\\degr$. It gives an estimate of jets age as $\\sim$ 1000 years. A simple kinematic model of SS433 (Abell {\\&} Margon 1979) with collimated jets was verified as a whole by many observational data (Margon 1984). SS433 is still the unique object in the Galaxy, in which the relativistic opposing jets appear in Doppler-shifted optical and X-ray emissions, i.e. beside the relativistic radio-emitting electron-positron plasma a baryonic matter with high atomic numbers moves in the SS433 jets (Kotani et al. 1996, Marshall et al. 2002). The major contribution to the study of temporal and spectral properties of microquasars was made by long-term program of variable sources monitoring at two frequencies of 2.25 and 8.3 GHz with GBI (Green-Bank Interferometer, NRAO). The publicly available data totally include about 16000 measurements of flux density of SS433. Fig.\\ref{gbi} shows the light curve of SS433 in 1979--2000 which cover all its GBI-observations at a frequency of 2.25~GHz. The results of these studies were partly published by Johnston et al. (1984) and Fiedler et al. (1987). From recent interesting results of the SS433 study noteworthy is the circular polarization of its emission in cm range detected by Fender et al. (2000) and a sign change of the Stokes parameter V detected by McCormick et al. (2003) in the 1--9~GHz range. It might be related either to the presence of a gyro-synchrotron radiation emitted by relatively low-energy electrons or to the change in a configuration of collimated magnetic field in the jets. Space observatory INTEGRAL first registered a variable hard X-ray emission (25--100~keV) (Cherepashchuk et al. 2003). It was shown that the emission level is influenced by the precession and orbital motions. These observations in high energy range confirm indirectly once more a presence of a black hole in this XB. In radio band a quasisteady synchrotron emission of SS433 with spectrum S$_{\\nu }$~$\\sim $~S$_{0}\\nu ^{\\alpha }$ is superimposed by a non-thermal flares, when the flux density can exceed the quiescent level 10 times. The SS433 radio flares are not rare, although the periods of relative quiescence can be as long as 100--200~days. At the flares onset there are evidences of synchrotron source opacity (a flat spectrum) and with further gradual decay of the flare the spectrum becomes steeper. The maximum flux of a flare comes later at lower frequencies as was shown for SS433 by Fiedler et al. (1987), Vermeulen et al. (1993a,b), Bursov {\\&} Trushkin (1995); for Cyg X-3 see Waltman et al. (1994, 1995, 1996) and for GRS~1915+105 --- Trushkin et al. (2001) and Fender et al. (2002). It corresponds to evolution of an adiabatically expanding cloud of relativistic emitting electrons. Shklovskij (1960) and van der Laan (1966) were the first to formulate basic equations describing this evolution. \\begin{figure*} \\centerline{\\vbox{ \\psfig{figure=1909gbi.eps,width=16cm,angle=-90} }} \\caption{% The SS433 light curves at a frequency of 2.25 GHz by data of GBI (NRAO/NASA).} \\label{gbi} \\end{figure*} Different modifications of this model necessary for observations fitting include a hollow conic geometry of jets, an account for synchrotron losses or reverse Compton scattering, a dynamic motion outwards a thermally absorbing envelope. Computations of flares based on a model of twin conic jet show a remarkable coincidence with observations for Cir\\,X-1 (Garcia 1995), SS433 (Hjellming {\\&} Johnston 1988), LSI+61$\\degr$\\,303 (Paredes et al. 1991), {\\mbox Cyg\\,X-3} (Marscher at al. 1975, Marti et al. 1992) and SAXJ1819-254 (Hjellming et al. 2000). Based on data of ten SS433 flares in 1986--87, Trushkin (1989) showed that on average they have close frequency power relations for the maximum flare flux and the time of this maximum occurrence: $$\\Delta S_{m}{\\rm(Jy)}=1.3 \\nu^{-0.4\\pm 0.15},$$ $$\\Delta t_{m}{\\rm (days)}=5\\nu^{-0.4\\pm 0.1},$$ where $\\nu$ is in GHz. A gradual flare decay follows one of two time laws. An original model of Shklovskij -- van der Laan provides the power law. In many microquasars there was actually detected the radio emission decay according to the law $$S_\\nu = S_o\\,t^{-2p}\\nu^{\\alpha},$$ where $p\\approx 2$ is an index of a power electron distribution that is related to the spectral index $$\\alpha = (1-p)/2\\approx-0.5$$ for an optically thin source. It is in excellent agreement with a possible diffusive mechanism of relativistic electrons acceleration in shock waves. At a compression index in a strong shock wave $\\rho\\le4$ the spectral index of these electrons is $$p = (\\rho+2)/(\\rho-1) = 2.$$ \\begin{table*} \\caption { A flux density sensitivity of the ``Northern Sector'' of RATAN-600} \\begin{center} \\medskip \\begin{tabular}{l|l|l|l|l|l|l|l} \\noalign{\\smallskip} \\hline Wavelength (cm) &31.2 &13.0 & ~7.6 &~6.2 & ~3.9 & ~2.7& 1.38 \\\\ \\hline Frequency (GHz) & 0.96 &~2.3 & ~3.9 &~4.9 & ~7.7 & 11.2& 21.7 \\\\ \\hline $\\Delta$S(mJy) & 40 & 15 & ~3 & ~3 & 8 & 10 & 20 \\\\ \\hline \\end{tabular} \\end{center} \\end{table*} Other flares and fluxes of plasmons in jets were decaying according to the exponential law $$S_{\\nu }=S_{0}e^{-t/ \\tau }$$ (for SS433 see Jowett \\& Spencer 1995). A diminution of the characteristic time $\\tau $ with frequency was often but not always observed. A relation $$\\tau {\\rm(days)}=11.5~\\nu ^{-0.29\\pm 0.03}$$ where $\\nu$ is in GHz was observed in Cyg X-3 (Trushkin 1998). In Cyg X-3 a transit from the exponential law to the power one was detected several times (Hjellming et al. 1974, Marscher et al. 1975, Bursov {\\&} Trushkin 1995, Trushkin 1998) what is interpreted as a transit from a mode with a dominated radiative losses to a mode of adiabatic expansion. Spencer (1996) calculated the energy of the jet components based on the condition for minimum energy of relativistic particles and of magnetic field what is fulfilled at an equipartition of energy of these components. He showed that a considerable not to say prevailing part of energy released during flares belonged to radio emitting particles and magnetic field of jets. Below we discuss a few large monitoring programs carried out at the Northern Sector of RATAN-600 having resulted in the SS433 light curves at 4--6 frequencies. ", "conclusions": "This paper compile all observational data over many sets from December 1986 to January 2004 containing 940 observations of SS433 at two--six frequencies simultaneously. In all more than 4500 measurements of flux density at frequencies from 960 to 21700~MHz are presented. The average spectral index by all data is equal to --0.60~$\\pm $~0.14 with an average measurement error of 0.09 at an average flux density of 1.5~Jy at 960~MHz. In the indicated frequency range there were no cases of the spectrum inversion, when the spectral index would be positive. Thus, if there exists such a stage of the SS433 spectrum evolution, it is essentially shorter than one day not to get in such a moment during almost 1000 observations. It was demonstrated that there are some earlier unknown regularities of the SS433 activity in the measured light curves. By the data on many flares the delay of flare maximum towards lower observational frequencies was detected. On the other side, the value of this maximum falls with frequency according to a power law. The flares fading with time also follows a power law. The indices of these relations change from flare to flare, but median values of indices coincide astonishingly well. They are equal to --0.4~$\\pm $~0.1. It is evidently an indication that the integral emission of the SS433 jets is due to similar properties of synchrotron radiation of the blobs--plasmons moving separately. In the end it is rigidly related with processes inside this XR and jets on the whole. From this point of view the models with inner shock waves propagating and increasing radio emission along the jet could be more preferable than the models in which the evolutions of separate radio components, blobs, are independent from each other. The accumulated data of the flux density measurements are presented as publicly-accessible Web-programs for constructing the SS433 spectra in the Home Page of the data base CATS: http://cats.sao.ru/cgi-bin/ss433.cgi. The complete data of all measurements carried out with RATAN-600 are presented in Table~3." }, "0403/astro-ph0403084_arXiv.txt": { "abstract": "We present the results of a wide-area mapping of the far-infrared continuum emission toward the Orion complex by using a Japanese balloon-borne telescope. The 155-$\\mu$m continuum emission was detected over a region of 1.5 deg$^2$ around the KL nebula with $\\timeform{3'}$ resolution similar to that of the IRAS 100-$\\mu$m map. Assuming a single-temperature model of the thermal equilibrium dust, maps of the temperature and the optical depth were derived from the 155 $\\mu$m intensity and the IRAS 100 $\\mu$m intensity. The derived dust temperature is 5 -- 15 K lower and the derived dust optical thickness were derived from the 155-$\\mu$m intensity and the IRAS 100-$\\mu$m intensity. The derived dust temperature is 5 -- 15 K lower and the derived dust optical depth is 5 -- 300 times larger than those derived from the IRAS 60 and 100-$\\mu$m intensities due to the significant contribution of the statistically heated very small grains to the IRAS 60-$\\mu$m intensity. The optical-thickness distribution shows a filamentary dust ridge that has a $\\timeform{1D.5}$ extent in the north -- south direction and well resembles the Integral-Shaped Filament (ISF) molecular gas distribution. The gas-to-dust ratio derived from the CO molecular gas distribution along the ISF is in the range 30 -- 200, which may be interpreted as being an effect of CO depletion due to the photodissociation and/or the freezing on dust grains. ", "introduction": "The Orion region at a distance of 470 pc \\citep{Brown1994}, is the nearest active star-forming region containing massive stars. It contains a massive molecular cloud complex, bright H\\emissiontype{II} regions, and many young stellar objects. Many studies of this region as a representative case of a massive-star-forming region have been made and have provided insight into the massive-star-formation process (see reviews by \\cite{Goudis1982, Genzel1989, O'Dell2001}). Even now, it is the most important region for investigating the massive-star-formation phenomena. M 42 is the brightest H\\emissiontype{II} region in this region, and is excited by the Trapezium star cluster. A dense molecular cloud complex lies behind the H\\emissiontype{II} region. In the densest part of this molecular could (OMC-1) are the Orion KL and BN objects \\citep{KL, BN}. They are embedded deeply in OMC-1, and energize the surrounding molecular gas of OMC-1. Two other dense molecular clouds, OMC-2 and -3, lie along the dense molecular ridge north of Orion KL. This dense ridge has a filamentary structure, $\\timeform{1.5D}$ long. From the apparent shape seen in the ${}^{13}$CO high-resolution map by \\citet{Bally1987}, it was named the Integral-Shaped Filament (hereafter ISF). The cloud cores of OMC-1, -2, and -3 lie on the dense filamentary feature. At the northernmost position of ISF, there is a H\\emissiontype{II} region, NGC 1977, and the south of it, there are small H\\emissiontype{II} regions, L 1641-N and NGC 1999. A H\\emissiontype{II} region, M42 containing the Orion KL, lies on the middle of the ISF. The ISF has so far been observed and identified in various molecular lines of CO (J = 2--1) by \\citet{Sakamoto1994}, ${}^{13}$CO (J = 1--0) by \\citet{Nagahama1998}, C${}^{18}$O (J = 1--0) by \\citet{Dutrey1991}, NH${}_3$ by \\citet{Cesanori1994}, and CS by \\citet{Tatematsu1993}, and in the neutral atomic carbon [C\\emissiontype{I}] line by \\citet{Ikeda2002}. The continuum emission in the far-infrared and sub-millimeter wavelength regions is radiated by interstellar dust grains that are mixed with the molecular gas. Thanks to recent progress in instrumental techniques in this wavelength region, relatively wide areas of the sky have been mapped with ground-based telescopes \\citep{Chini1997, Lis1998, Johnstone1999} and with balloon-borne telescopes \\citep{Ristorcelli1998, Mookerjea2000-a, Dupac2001}. Previous investigations successfully revealed the dust distribution of this region, but covered only a part of the ISF-OMC-1 and its vicinity, OMC-2, and OMC-3. \\citet{Mookerjea2000-a} mapped this region at 138 and 205 $\\mu$m with a spatial resolution of $\\timeform{1.5'}$ using a one-meter balloon-borne telescope, and showed that a dust ridge extends from Orion KL to the north, tracing the molecular ridge. \\citet{Ristorcelli1998, Dupac2001} obtained wider-area maps at 200, 260, 360, and 580 $\\mu$m with the large balloon-borne telescope, PRONAOUS. The observed area was from Orion KL to the north. Their observations revealed several cold condensations to the west of OMC-1, which may be pre-collapse phase cloud cores. \\citet{Johnstone1999} mapped the southern ridge with SCUBA of JCMT, and showed that the dust ridge really extends along the ISF at least up to $\\timeform{30'}$ from Orion KL. However, the observed area ($\\timeform{50'} \\ \\times \\ \\timeform{10'}$) did not cover the whole of the ISF. It is of great interest to map the interstellar dust distribution over the whole region of the ISF. As is well known, IRAS completely mapped this region at four infrared bands of 12, 25, 60, and 100 $\\mu$mm. By using the IRAS 60 and 100 $\\mu$m maps, \\citet{Bally1991} derived the maps of dust temperature and the dust mass. However, they could not identify the ISF-like structure in the derived maps. The main reason for this is the unsuitability of the IRAS 60 $\\mu$m band for the study of a cold dust region like the ISF, because in this band, the emission from stochastically-heated, very small grains significantly contaminates the thermal emission from the large grains at a steady state temperature (e.g. \\cite{Nagata2002}). Moreover, as is mentioned later, the IRAS 60 and 100 $\\mu$m band signals are saturated at and near the Orion KL position and have a spurious pattern due to the insufficient baffle design. COBE/DIRBE had achieved highly reliable, well-calibrated intensity maps at various infrared bands for the whole sky. The 100, 140, and 240 $\\mu$m intensity maps of DIRBE can be used for deriving the dust temperature and mass distribution. However, the spatial resolution of $\\timeform{0.7D}$ of the DIRBE maps is insufficient for resolving the dust distribution in the Orion region \\citep{Wall1996}. The aim of the present work is to observe a very large area covering the whole ISF at wavelengths between 100 and 200 $\\mu$m, where the interstellar dust (large grains) has its peak emission, with a spatial resolution similar to that of IRAS 100 $\\mu$m band in order to reveal the dust temperature and mass distribution and, consequently, the energy source of individual parts of molecular clouds in this region. The instrumentation and observation are described in section 2; the data reduction, analysis, and calibration are explained in section 3. The intensity map and the temperature distribution are presented in section 4. In section 5, the column-density distribution of the dust grain is derived, and the heating sources for this region are discussed. ", "conclusions": "This paper presents the result of a wide-area mapping of the 155 $\\mu$m continuum emission in the Orion molecular cloud complex. The Japanese balloon-borne telescope (FIRBE) surveyed a sky area over 50 deg$^2$ and detected far-infrared emission over the region of 1.5 deg$^2$ around the KL nebula with $\\timeform{3'}$ resolution. The peak emission at Orion KL has a flux value of 42.5 $\\pm$ 4.8 kJy in the $\\phi \\ \\timeform{3'}$ beam. The far-infrared distribution of intense region is similar that of the IRAS 100 $\\mu$m image. The dust temperature and optical depth of the thermal-equilibrium large grain are estimated by applying a single-temperature dust model modified with a $\\lambda^{-2}$ emissivity law to the IRAS 100 and FIRBE 155 $\\mu$m intensities. They range from 15 to 30 K and from 0.15 to 0.0015, respectively. Due to the significant contribution of the statistically heated very small grain in the IRAS 60 $\\mu$m intensity, the dust temperature T(60/100) derived from the IRAS 60 and IRAS 100 $\\mu$m intensities is 5 -- 15 K higher than the thermal-equilibrium dust temperature. As a result, the dust optical depth at 100 $\\mu$m, derived from T(60/100) has a 5 -- 300 times smaller value. Toward the H\\emissiontype{II} region, M 42, the IRAS 60 $\\mu$m and also IRAS 25 $\\mu$m intensities are relatively stronger than those outside of the H\\emissiontype{II} region. This tendency in the H\\emissiontype{II} region may be attributed to the result of the statistically heating of very small grains by double photon incidence \\citep{Okumura1999}. The region where the dust temperature is greater than 20 K around M 42 has an extent similar to that of the optical region. This fact suggests that the far-infrared emission around M 42 is mainly provided by the Trapezium Cluster. The optical-depth distribution shows a filamentary dust ridge that has a $\\timeform{1.5D}$ extent in the north -- south direction. The shape of this dense dust ridge resembles the Integral-Shaped Filament (ISF) molecular gas distribution. The gas-to-dust ratio derived from the CO molecular gas distribution along the ISF shows 100 -- 200 around the M 42 ionized region on the ISF. However, the gas-to-dust ratio decreases in other regions of the ISF, which may be interpreted as being an effect of CO depletion due to the photodissociation and/or freezing on dust grains. \\bigskip We would like to acknowledge Dr. T. Nagahama and Dr. Y. Fukui for allowing us to use the ${}^{13}$CO data and Dr. X. Dupac for providing us with the PRONAOUS data. We also express our special thanks to the technical staff of Tata Institute of Fundamental Research and National Balloon Facility of India. This work was supported by the Grant-in Aid of the Japan Society for the Promotion of Science (10147102) and by the Research Fellowship of Japan Society for the Promotion of Science. \\clearpage \\begin{figure} \\begin{center} \\FigureFile(80mm,80mm){Figure1.eps} \\end{center} \\caption{ The thick curve shows the relative systematic spectral response of the FIRBE 155 $\\mu$m band. The thin-curve and the broken-curve are the spectral responses of the IRAS 100 $\\mu$m band and the COBE/DIRBE 140 $\\mu$m band, respectively (IRAS and COBE/DIRBE Explanatory Supplement). All curves are normalized at their peaks. } \\end{figure} \\clearpage \\begin{figure} \\begin{center} \\FigureFile(80mm,80mm){Figure2.eps} \\end{center} \\caption{ Example of raw detector signals of eight pixels in one column of the 4 by 8 pixel array at crossing Orion KL. The horizontal axis shows the azimuth offset relative to the peak position that corresponds to Orion KL. } \\end{figure} \\clearpage \\begin{figure} \\begin{center} \\FigureFile(160mm,160mm){Figure3.eps} \\end{center} \\caption{ (a) 155 $\\mu$m continuum intensity map observed by the FIRBE telescope. The spatial resolution is $\\timeform{3'}$ and the intensity unit is MJy sr$^{-1}$. The lowest contour levels are 352 (=2.6 $\\sigma$), 488 (= 3.6$\\sigma$), 600, 700, 800, 900, 1000, 1300, 1000 interval up to 10000, 10000 interval up to peak intensity of 71080 MJy sr$^{-1}$ corresponding to the flux of 42.5 kJy in $\\phi \\ \\timeform{3'}$ beam. (b) The IRAS 100 $\\mu$m intensity map of the same area after correction for the DIRBE -- IRAS calibration scale difference (multiplied by 0.72). The contour levels are 70, 100, 150, 100 interval from 200 up to 1000, and 1000 interval up to 10000. The IRAS 100 $\\mu$m intensity around Orion KL is saturated and the sprious patterns appear at an angular distance around $\\timeform{1D}$ away from Orion KL. } \\end{figure} \\clearpage \\begin{figure} \\begin{center} \\FigureFile(160mm,160mm){Figure4.eps} \\end{center} \\caption{ (a) Dust temperature ($T_{\\mathrm{LG}}$) map derived from the 155 $\\mu$m intensity of the present work and the IRAS 100 $\\mu$m intensity. The contours indicate 16, 20, 24, and 28 K. (b) Same as (a) from the IRAS 60 and 100 $\\mu$m intensities. The contours indicate at intervals of 4 K from 24 to 40 K. (c) Red-band image of the Digital Sky Survey overlaid on the contour map of (b). The white-colored region represents the H\\emissiontype{II} region determined from the 330 MHz radio continuum map by \\citet{Subrahmanyan2001}. } \\end{figure} \\clearpage \\begin{figure} \\begin{center} \\FigureFile(100mm,100mm){Figure5.eps} \\end{center} \\caption{ Schematic diagram for defining the three regions around M 42 superposed on the red-band image. The gray central rectangular region indicates the region where the IRAS 100 $\\mu$m band is saturated. } \\end{figure} \\clearpage \\begin{figure} \\begin{center} \\FigureFile(160mm,160mm){Figure6.eps} \\end{center} \\caption{ (a)Gas column density map derived from the dust optical depth and a gas-to-dust mass ratio of 100. Logarithmic contours are shown along the ISF at the interval of 0.2 from 22.5 up to 23.5. (b) ${}^{13}$CO (J = 1--0) total integrated intensity map with $\\timeform{3'}$ spatial resolution \\citep{Nagahama1998}, where the velocity range is from 0 to 15 K km s$^{-1}$, is overlaid on the color map described in figure 6a. Contour levels are at 3.0, 5.0, 7.0, and at the intervals of 5 K km s$^{-1}$ from 10.0 K km s$^{-1}$. } \\end{figure} \\clearpage \\begin{figure} \\begin{center} \\FigureFile(160mm,160mm){Figure7.eps} \\end{center} \\caption{ (a) The gray area indicates the area for which the dust mass was derived as in (b). The contour is the ${}^{13}$CO integrated intensity map. The vertical axis is the declination offset relative to the OMC-1 position. (b) Mean dust mass distribution along the ISF ridge, averaged in each $\\timeform{3'}$ bin in declination. (c) Same as (b) for the gas mass. (d) Gas-to-dust distribution. } \\end{figure} \\clearpage \\begin{table} \\caption{Intensity ratio of the `High' Temperature Region to the surrounding reference region.} \\begin{center} \\begin{tabular}{lccccc} \\hline \\hline & \\multicolumn{4}{c}{\\textit{IRAS}} & \\textit{FIRBE} \\\\ \\cline{2-6} Band & 12 $\\mu$m & 25 $\\mu$m & 60 $\\mu$m & 100 $\\mu$m & 155 $\\mu$m \\\\ \\hline Intensity ratio & 1.30 $\\pm$ 0.20 & 1.72 $\\pm$ 0.42 & 1.71 $\\pm$ 0.36 & 1.26 $\\pm$ 0.21 & 1.25 $\\pm$ 0.25 \\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\clearpage" }, "0403/astro-ph0403617_arXiv.txt": { "abstract": "{We report a flare on the M9 dwarf DENIS 104814.7-395606.1, whose mass places it directly at the hydrogen burning limit. The event was observed in a spectral sequence during 1.3 hours. Line shifts to bluer wavelengths were detected in $\\mathrm{H_{\\alpha}}$, $\\mathrm{H_{\\beta}}$, and in the Na\\,{\\sc i} D lines, indicating mass motions. In addition we detect a flux enhancement on the blue side of the two Balmer lines in the last spectrum of our series. We interpret this as rising gas cloud with a projected velocity of about 100 $\\mathrm{km s^{-1}}$ which may lead to mass ejection. The higher Balmer lines $\\mathrm{H_{\\gamma}}$ to $\\mathrm{H_{8}}$ are not seen due to our instrumental setup, but in the last spectrum there is strong evidence for $\\mathrm{H_{9}}$ being in emission. ", "introduction": "DENIS 104814.7-395606.1 (hereafter DENIS 1048-39) was discovered in the DEep Near Infrared Survey (DENIS), which covered the southern sky \\citep{Epchtein} in two near infrared bands (J and $\\mathrm{K_{s}}$) and one optical band (I). With this choice of bands DENIS is very sensitive to very low mass stars and brown dwarfs and thus excellently suited for searches for hitherto unknown low mass stellar or substellar objects in the solar vicinity. With a distance of only $4.6\\pm0.3$\\,pc \\citep{DENISinfr} DENIS 1048-39 is extremely close to the Sun. Classified as a M9 star, the lithium resonance line at \\mbox{6708 \\AA} could not be detected, therefore substantial lithium depletion must have taken place \\citep{DENISdelfosse}. Compared to LP~944-20, which is also classified as M9, DENIS~1048-39 should be older and more massive. Theoretical models place DENIS 1048-39 directly at the hydrogen burning boundary with an estimated mass of 0.075 up to 0.09 $\\mathrm{M_{Sun}}$ and an age of 1 - 2 Gyrs \\citep{DENISinfr} assuming a solar chemical composition. Thus DENIS 1048-39 may either be among the most massive brown dwarfs or among the least massive stars. Interestingly, the $\\mathrm{H_{\\alpha}}$-line was found to be variable \\citep{DENISdelfosse}, implying that DENIS 1048-39 exhibits activity despite its old age and low mass. There are only a few very late type objects showing strong and persistent H$_{\\alpha}$ emission. H$_{\\alpha}$ emission, a well established activity indicator, is not detected in the majority of the stars. Instead one observes a steep decline of the strength of H$_{\\alpha}$ emission for stars later than spectral type M7. This effect shows up in smaller equivalent widths (EW) of the H$_{\\alpha}$ line for these objects. However, since the H$_{\\alpha}$ line is seen against an increasingly faint photosphere for later objects, a better activity indicator is the ratio of the H$_{\\alpha}$ luminosity to the bolometric luminosity. This ratio was found to drop in only three subclasses (M8-L1) by one order of magnitude \\citep{Gizis}. The same authors also found that the activity of these late type objects is primarily related to temperature and shows only a secondary dependence on mass and age. Thus strong H$_{\\alpha}$ emission does not necessarily imply youth; on the contrary, strong H$_{\\alpha}$ emitters in the field are more likely to be old. Despite this decline in activity there are many reports of flare activity even among L dwarfs. For very late-type M dwarfs \\citet{Gizis} estimated a flaring time fraction of about 7 \\%, while \\citet{Liebert1} found a flare duty cycle of about 1 \\% for L dwarfs, suggesting that there must be some ongoing magnetic activity in these dwarfs despite the absence of a persistent chromosphere or corona. In this paper we report on a huge flare on DENIS 1048-39 detected in the H$_{\\alpha}$, H$_{\\beta}$, and sodium emission lines. In section 2 we describe the VLT data and their analysis, in section 3 we discuss the timing behaviour of the flare. ", "conclusions": "\\subsection{Emission geometry} The measured half widths of the Na\\,{\\sc i} D and the H$_{\\beta}$ lines always stay below 20 $\\mathrm{km\\,s^{-1}}$, i.e., they are smaller than the measured rotational velocity. The same applies to the two main components of the H$_{\\alpha}$ line. This suggests that the emission for the Balmer lines as well as for the Na\\,{\\sc i} D lines is confined to a restricted region on the star. Since the H$_{\\alpha}$ line consists of two components and there is evidence that the H$_{\\beta}$ line consists of two components as well, the star could host even two active regions producing the observed Balmer emission. Now the question arises what the nature of these two active regions may be. There are two main possibilities: a dynamic scenario or a static scenario. \\subsection{Static or dynamic scenario ?} The dynamic scenario invokes two active regions located on the surface of the star and mass motions within these active regions. The Balmer emission comes from the chromosphere, nevertheless very close to the surface of the star. At least in one of the two regions a brightening in the Balmer lines takes place on a timescale of about 1.5 hours. H$_{\\alpha}$ and H$_{\\beta}$ emitting material moves towards the observer and causes the blue shift. The static scenario interprets the two regions as emitting gas clouds corotating in some distance to the stars surface as was first proposed for the K0 star AB~Dor by \\citet{Cameron}. Since for one component the line shift becomes bluer this one must be rising while the other one must be about to set behind the star since its line shift is moving towards the blue. The blue shifted component must then be captured during its rise above the horizon of the star to account for the flux increase. This latter static scenario can be excluded because of the rapid rotation of the stars as follows: For an estimated radius of $\\mathrm{R_{\\star}} = 0.1 \\mathrm{R_{\\odot}}$ the maximal rotation period of DENIS 1048-39 is 4.9 hours. Therefore in the four consecutive spectra lasting together 80 minutes the star completes about a fourth of its rotation. If the emitting gas is confined in a corotating cloud, the measured velocity shift should exceed the star's rotational velocity in at least one spectrum for orbits near the equatorial plane. Moreover the flux increase cannot be explained by a cloud rotating into view because the star rotates too fast for this interpretation. A more natural explanation is therefore the scenario of active regions with mass motions at the surface of the star that brighten during the spectral series. \\subsection{Blue-shifted Balmer emission} Let us now consider the broad emission features bluewards of the two Balmer lines seen only in the fourth spectrum. Again a dynamic or static scenario may be considered. Let us first assume a static interpretation with a corotating cloud. Since its radial velocity of about 100 $\\mathrm{km\\,s^{-1}}$ is too high to be interpreted by a region on or near the surface of the star, the emission would have to come from a cloud at some distance to the star. Such a cloud must be confined then by magnetic loops above the surface of the star and has rotated just into the field of view. If one assumes a stellar inclination close to $90^{\\circ}$ and a cloud in the equatorial plane, one can compute its distance $R$ from the rotation axis with $R=\\frac{\\mathrm{v} P}{2\\pi}$ with $\\mathrm{v}$ denoting the radial velocity and $P$ the rotation period of DENIS~1048-39. We find a distance of $R=4 \\mathrm{R_{\\star}}$ with the measured radial velocity of 100 $\\mathrm{km\\,s^{-1}}$, which is below the Keplerian corotation radius of $6 \\mathrm{R_{\\star}}$ like the clouds found on AB~Dor in $\\mathrm{H_{\\alpha}}$ \\citep{Donati} and other chromospheric lines \\citep{Brandt}. But since the star rotates fast, the feature should be visible in more than one spectrum since for its distance of four stellar radii it can be shadowed by the star only about 10 \\% of the time, i.\\,e. about 30 minutes, thus again a dynamic interpretation appears more likely. In the dynamical interpretation the line shift is explained in terms of material ejected by the star. Since the emission in the main body of the $\\mathrm{H_{\\alpha}}$ brightens substantially in the same spectrum it is suggestive that the broad emission feature is connected to the active region producing the $\\mathrm{H_{\\alpha}}$ line arguing further in favor of a dynamic scenario. In the context of interpreting the broad emission feature as the signature of a rising cloud the question remains unsettled whether this rise leads to mass ejection since we do not know the longitude of the cloud. Assuming again an radius of $0.1 \\mathrm{R_{\\odot}}$ we find an escape velocity of about 550 $\\mathrm{km\\,s^{-1}}$ for a mass between 0.075 and 0.09 $\\mathrm{M_{\\odot}}$ while the projected velocity of the cloud is about 100 $\\mathrm{km\\,s^{-1}}$. Let us now consider the width of the emission feature. This width can be due to temperature and turbulent broadening if one thinks of a confined gas cloud rising towards the observer. A first estimate of the temperature of a $\\mathrm{H_{\\alpha}}$ emitting gas is about 10\\,000 K leading to a thermal broadening of 9 $\\mathrm{km\\ s^{-1}}$, far less than the observed total line broadening of 50 $\\mathrm{km\\ s^{-1}}$. In this scenario with a rising cloud one can estimate the height of the cloud after the twenty minutes exposure. Assuming a constant cloud velocity of the cloud and an ejection start right at the beginning of the exposure leads to a cloud height of 1.7 $\\mathrm{R_{\\star}}$. It is probable that the cloud has been decelerated during such a rise. Therefore a second interpretation of the line width is that different velocities of the cloud during deceleration are integrated over the exposure time. Moreover, the cloud may consist of more than one component with different velocities. Besides these uncertainties in the event geometry there is no doubt that a dynamical interpretation of the spectra is needed. The last spectrum is then quite suggestively interpreted as the onset of a flare on DENIS~1048-39 as reported for AD~Leo by \\citet{Houdebine} who found a similar flux enhancement in the far blue wing of the $\\mathrm{H_{\\gamma}}$ line during a particular violent flare on AD~Leo. Since these authors found projected velocities of up to 5800 $\\mathrm{km\\,s^{-1}}$ this event was clearly associated with a mass ejection, while this question remains unsettled for DENIS~1048-39. \\subsection{Summary} In conclusion, we find at least two active regions on DENIS~1048-39 contributing to the bulk of the Balmer line flux. Mass motions directed towards the observer are found for the emission lines of the Balmer series as well as for the Na D lines. In the last spectrum of our observations the onset of a flare seems to take place, since substantial brightening and blueshifts can be seen in the lines. In addition in the $\\mathrm{H_{\\alpha}}$ and the $\\mathrm{H_{\\beta}}$ line there is a broad emission feature on the blue side of the line. This can be interpreted as a rising cloud. Since DENIS~1048-39 seems to be located directly at the hydrogen burning limit this flare gives evidence that such events may be more ubiquitous than previously assumed. It is consistent with X-ray detections of brown dwarfs \\citep{Mokler} and the X-ray flare event found on the similar late-type star LHS~2065 \\citep{LHS-2065}." }, "0403/hep-th0403234_arXiv.txt": { "abstract": "We discuss an inflation model, in which the inflation is driven by a single scalar field with exponential potential on a warped DGP brane. In contrast to the power law inflation in standard model, we find that the inflationary phase can exit spontaneously without any mechanism. The running of the index of scalar perturbation spectrum can take an enough large value to match the observation data, while other parameters are in a reasonable region. ", "introduction": "Astronomical observations in resent years, especially by Wilkinson Microwave Anisotropy Probe (WMAP) \\cite{wmap}, lead to a high precision era of cosmology. All the results imply there exists an accelerating moment (inflationary phase) in the very early universe. The angle power spectrum of cosmological microwave background (CMB) provides some evidences that the universe is almost spatially flat and that large scale structure is formed from a primordial spectrum of adiabatic, normal and nearly Harrison-Zeldovich (scale invariant) density perturbations. This can be explained by the simplest model of inflation~\\cite{liddlelyth}. Despite the great success of the big bang standard model of cosmology together with inflation, there are still several serious problems in our present scenario of cosmology. The cosmological constant (dark energy), unexpected low power spectrum at large scales, egregious running (to the common inflationary models) of power spectrum index are distinctive ones. The latter two might relate to the very high energy physics. Although many approaches have been made in the literature, fairly speaking, these problems still stick on. In view of achievements and shortcomings of inflationary scenario in Einstein gravity, it is necessary to further deepen our understanding of the inflationary scenario from a theoretical perspective. In nonperturbative string/M theories there are topological solitons, called branes, in $10$ or $11$ dimensional spacetime. In the Horava-Witten model \\cite{horava}, gauge fields of the standard model are confined on two 9-branes located at the end points of an $S^1/Z_2$ orbifold, i.e., a circle folded on itself across a diameter. Inspired by the Horava-Witten model, the idea that our universe is a 3-brane embedded in a higher dimensional spacetime has received a great deal of attention in recent years~\\cite{braneworld}. In this brane world scenario, the standard model particles are confined on the 3-brane, while the gravitation can propagate in the whole space. In this picture, the Friedmann equation of the brane universe gets modified~\\cite{branecos}, compared to the one for the standard model. The chaotic inflation model on the RSII brane has been studied in~\\cite{chaotic}. The inflation model in the brane world scenario with a Gauss-Bonnet term in the bulk has also been discussed recently by Lidsey and Nunes~\\cite{LN}. In the RSII model~\\cite{RSII}, a brane with positive tension is embedded in five dimensional anti de Sitter space. Due to the warped effect of bulk geometry, general relativity on the brane can be recovered in low energies, while gravity on the brane is five dimensional in high energy limit. Compared to the RSII model, the brane world model proposed by Dvali, Gabadadze and Porrati (DGP)~\\cite{dgpmodel} is also very interesting. In the DGP model, the bulk is a flat Minkowski spacetime, but a reduced gravity term appears on the brane without tension. In this model, gravity appears 4-dimensional at short distances but is altered at distance large compared to some freely adjustable crossover scale $r_0$ through the slow evaporation of the graviton off our 4-dimensional brane world universe into an unseen, yet large, fifth dimension. The late-time acceleration is driven by the manifestation of the excruciatingly slow leakage of gravity off our four-dimensional world into an extra dimension. This offers an alternative explanation for the current acceleration of the universe. In the DGP model, the gravitational behaviors on the brane are commanded by the competition between the 5-dimensional curvature scalar ${}^{(5)}R$ in the bulk and the 4-dimensional curvature scalar $R$ on the brane. At short distances the 4-dimensional curvature scalar $R$ dominates and ensures that gravity looks 4-dimensional. At large distances the 5-dimensional curvature scalar ${}^{(5)}R$ takes over and gravity spreads into extra dimension. As a result, Newton-like force law becomes 5-dimensional one. Thus, gravity begins weaker at cosmic distances. So it is natural that such a dramatic modification affects the expansion velocity of the Universe. In fact the DGP model has been applied to cosmology immediately after DGP putting forward their model \\cite{dgpcosmology}. In this paper, we will study an inflation model in a brane world scenario, which combines the RSII model and DGP model. That is, an induced curvature term will also appear on the brane in the RSII model. We call the brane as warped DGP brane. In this model, inflation of the universe is driven by a single scalar field with exponential potential on the brane. In contrast to the power law inflation in standard model, we find that the inflationary phase exits spontaneously. The running of the index of the scalar perturbation spectrum can take negative values, which content the high precision observations by WMAP. At the same time other parameters are in a reasonable region. The organization of this paper is as follows. In the next section, we present the Friedmann equation on the warped DGP brane. In Secs.~III and IV the inflation model is introduced. We analyze the parameter space allowed by the observation data in Sec. V. The paper ends in Sec.~VI with conclusions and discussions. ", "conclusions": "Various inflationary models have been proposed since Guth's seminal work~\\cite{Guth}. Going with the observations of high precision, we realize that a successful inflation model must at least possess the following properties: 1) a sufficient large number of e-folds, 2) a near Harrison-Zeldovich (scale invariant) spectrum, 3) a negative running of spectrum index. The model we discussed in this paper contents these. In particular, it is worth noting that in this inflation model based on the warped DGP brane world scenario, even for an exponential potential, the inflationary phase can exit naturally, and within a reasonable parameter region, the model can give us a negative running of the scalar spectrum index. By choosing more appropriate parameters in the model, we may obtain much better consistence with the current observation data. These features are quite attractive. Other properties of this model deserves further study. For example, it would be interesting to investigate carefully the relations of the spectrum index and its running to the comoving wave number $k$. In addition, in the warped DGP brane model, it is certainly of interest to further construct inflation models with scalar spectrum index larger than one at larger scales. {\\bf Note added:} While we were finishing this paper, a paper~\\cite{new} appears on the net, which also discusses an chaotic inflation model on a brane with induced gravity. More recently, two related and interesting papers occur on the net. In \\cite{BMW} the authors calculate the amplitude of gravitational waves from brane-world inflation with induced gravity, while the authors of \\cite{BW} study the induced gravity with a non-minimally coupled scalar field on the brane. {\\bf Acknowledgments:} This work was supported in part by a grant from Chinese academy of sciences, a grant No. 10325525 from NSFC, and by the ministry of science and technology of China under grant No. TG1999075401." }, "0403/astro-ph0403329_arXiv.txt": { "abstract": "{ Three-dimensional hydrodynamic calculations are performed in order to investigate mass transfer in a close binary system, in which one component undergoes mass loss through a wind. The mass ratio is assumed to be unity. The radius of the mass-losing star is taken to be about a quarter of the separation between the two stars. Calculations are performed for gases with a ratio of specific heats $\\gamma=1.01$ and $5/3$. Mass loss is assumed to be thermally driven so that the other parameter is the sound speed of the gas on the mass-losing star. Here, we focus our attention on two features: flow patterns and mass accretion ratio, which we define as the ratio of the mass accretion rate onto the companion, $\\dot{M}_{acc}$, to the mass loss rate from the mass-losing primary star, $\\dot{M}_{loss}$. We characterize the flow by the mean normal velocity of wind on the critical Roche surface of the mass-losing star, $V_R$. When $V_R<0.4 A\\Omega$, where $A$ and $\\Omega$ are the separation between the two stars and the angular orbital frequency of the binary, respectively, we obtain Roche-lobe over-flow (RLOF), while for $V_R>0.7 A\\Omega$ we observe wind accretion. We find very complex flow patterns in between these two extreme cases. We derive an empirical formula of the mass accretion ratio as $0.18 \\times 10^{-0.75V_R/A\\Omega}$ ~in the low velocity regime and $0.05\\,(V_R/A\\Omega)^{-4}$ in the high velocity regime. ", "introduction": "Wind accretion plays an essential role especially in the evolution of detached binary systems such as symbiotic stars, precursor of peculiar red giants, $\\zeta$ Auriga stars and massive X-ray binaries. In the majority of symbiotic stars, the system contains a hot component - believed to be a mass-gaining white dwarf - and a cool component, the mass-losing red giant (e.g., Miko{\\l}ajewska \\cite{Miko} for a review). Recent studies indicate that many of the red giant components in symbiotic stars do not fill their Roche lobe, i.e., $R_{\\rm RG} \\la \\ell_1 /2$, where $R_{\\rm RG}$ is the radius of the red giant component and $\\ell_1$ is the distance of the inner Lagrangian point from the centre of the red giant (e.g., M\\\"urset \\& Schmid \\cite{1999A&AS..137..473M}). This possibly discloses that symbiotic stars are - with probably only the exception of T CrB (e.g., Anupama \\& Miko{\\l}ajewska \\cite{1999A&A...344..177A}) - well detached binary systems. Thus, mass transfer in symbiotic stars seems to be driven by wind, and wind mass loss is therefore a key ingredient for triggering symbiotic activity on a white dwarf companion. On the other hand, the mass accretion rates expected in the Bondi-Hoyle picture (Bondi \\& Hoyle \\cite{1944MNRAS.104..273B}) for such detached binary systems are usually much lower than those required to maintain the symbiotic activity of the systems. For example, Nussbaumer (\\cite{Nuss}) pointed out that in most symbiotic stars the typical rates of wind mass loss red giant are in the range $1 \\times 10^{-6} M_\\odot$ yr$^{-1}$ to $1 \\times 10^{-8} M_\\odot$ yr$^{-1}$ and that the expected accretion efficiency of wind capture is probably no better than one percent. Hence, the accretion rates associated with most symbiotic white dwarfs are probably in the range of $1 \\times 10^{-8} M_\\odot$ yr$^{-1}$ to $1 \\times 10^{-10} M_\\odot$ yr$^{-1}$, which are too low to power the typical luminosities of symbiotics ($\\ga 1000 L_\\odot$) unless they are all in a late phase of hydrogen shell-flash (novae) on a white dwarf (e.g., Sion \\& Ready \\cite{1992PASP..104...87S}; Sion \\& Starrfield \\cite{1994ApJ...421..261S}). It will nevertheless be shown in the present paper that the mass accretion rate is smaller than that expected by a simple Bondi-Hoyle picture. It has recently been suggested that some of the mass-accreting white dwarfs in symbiotic stars are the progenitors of Type Ia supernovae (e.g., Hachisu \\& Kato \\cite{2001ApJ...558..323H}). On the other hand, it has long been discussed that the efficiency, the ratio of the mass captured by the white dwarf to the mass lost by the red giant, is very small (say, a few per cent) and not enough to increase the white dwarf mass to the Chandrasekhar mass limit (e.g., Kenyon et al. \\cite{1993ApJ...407L..81K}). Wind velocities in cool red giants are typically a few times 10 km s$^{-1}$ and comparable to the orbital velocity in most symbiotic binary systems. Therefore, flow patterns are probably something between Roche lobe overflow (RLOF) and Bondi-Hoyle wind accretion kinds of flows. The efficiency of capture in symbiotics should therefore not be estimated by the Bondi-Hoyle picture but by direct simulations of intermediate, complicated flow patterns. Symbiotic stars are not the only systems in which wind accretion play a significant role. The most adopted model for the formation of peculiar red giants, Barium, CH and S stars, require a binary system containing an Asymptotic Giant Branch star (AGB) transferring mass via its stellar wind to a main sequence companion which becomes polluted in Carbon and s-process elements (see e.g. Boffin \\& Jorissen \\cite{BJ88}). The AGB then evolve into a white dwarf while the companion will appear as Carbon or s-process rich and, when on the giant branch, as a peculiar red giant. Boffin \\& Zacs (\\cite{BZ94}) have shown that an accretion efficiency of only a few percent is enough to explain present Barium stars. Objects related to these are the extrinsic S stars which also show symbiotic activity (e.g. Carquillat et al. \\cite{Carq98}) and bipolar planetary nebulae (e.g. Mastrodemos \\& Morris \\cite{MM98}). $\\zeta$ Aurigae systems are another kind of system where wind accretion plays a role. In these eclipsing double-lined binaries, a main sequence star has an accretion wake produced by the wind of its K supergiant companion. The eponymous system, $\\zeta$ Aur, is a binary with a 972 days orbital period and containing a 5.8 M$_\\odot$ K4 Ib and a 4.8 M$_\\odot$ K B5 V star (Bennett et al. \\cite{Benn96}). Finally, accretion wakes are also reported for high mass X-ray binaries (HMXB). These systems consist of a compact object, neutron star or black hole, orbiting a massive OB primary star which has a strong stellar wind. The X-ray emission is believed to be due to accretion of matter on the compact companion. There have been already many numerical studies of stellar wind in a close binary system. Biermann (\\cite{1971A&A....10..205B}) computed two-dimensional stellar wind using a characteristic method. Sorensen, Matsuda \\& Sakurai (\\cite{1975Ap&SS..33..465S}) performed two-dimensional finite difference calculation of stellar wind emitted from a Roche lobe filling secondary. They used the Fluid in Cell (FLIC) method with first order of accuracy and a Cartesian grid. They computed Roche lobe overflow (RLOF) as well. Sawada, Hachisu \\& Matsuda (\\cite{1984MNRAS.206..673S}) calculated two-dimensional stellar wind from a contact binary using a Beam-Warming time implicit finite difference scheme and a generalized curvilinear coordinate. Their main goal was to compute the angular momentum loss rate from the system, which is an important factor to define the evolution of the binary system. Sawada, Matsuda \\& Hachisu (\\cite{1986MNRAS.221..679S}) computed two-dimensional stellar wind from a semi-detached binary system using the Osher upwind scheme and a generalized curvilinear coordinate. They observed a transition from RLOF to stellar wind by increasing the wind speed on the mass-losing star. The present study is a three-dimensional version of their study. Matsuda, Inoue \\& Sawada, (\\cite{1987MNRAS.226..785M}) also conducted a similar study and discovered a flip-flop instability of the bow shock formed around a compact mass accreting object (see also Boffin \\& Anzer, \\cite{BA94}, and on the stability issue, Foglizzo \\& Ruffert, 1997 and 1999, and Pogolerov, Ohsugi \\& Matsuda, 2000). However, it was found that the phenomenon was characteristic to two-dimensional case (Matsuda et al., \\cite{1992MNRAS.255..183M}, Ruffert, \\cite{Ruffert96}). In the present work we perform three-dimensional calculation in which we do not observe the flip-flop instability. These simulations concerned the case of a compact object moving in a wind and were mostly aimed at estimating the validity of the Bondi-Hoyle accretion rate. Binary effects were partially simulated by including velocity or density gradient in the wind. There are a few simulations however which simulate in full mass transfer by wind in a binary system (Blondin et al. 1991, Theuns \\& Jorissen 1993, Blondin \\& Woo 1995, Theuns et al. 1996, Walder 1997, Mastrodemos \\& Morris 1998, Dumm et al. 2000, Boroson et al. 2001, Gawryszczak, Miko{\\l}ajewska, \\& R{\\' o}{\\. z}yczka 2002). Theuns \\& Jorissen (\\cite{TJ93}) and Theuns, Boffin \\& Jorissen (\\cite{TBJ96}, TBJ96 in the following) performed three-dimensional hydrodynamic simulations of a 3 M$_\\odot$ AGB transferring mass through its stellar wind to a 1.5 M$_\\odot$ main sequence companion, in order to test the wind accretion model for the origin of peculiar red giants. They performed simulations using a polytropic equation of state with $\\gamma =$1., 1.1 and 1.5 and found mass accretion ratios (that is, mass accretion rate/mass loss rate) of 1-2\\% for the $\\gamma =1.5$ case and 8 \\% for the other models, i.e. about ten times smaller than the theoretical Bondi-Hoyle estimate. They also observed that this value is dependent on resolution. Walder (1997) presented simulations of wind accretion in well separated binaries for three different cases: a HMXB, $\\zeta$ Aur and a barium star progenitor. He obtains mass accretion ratio (mass accretion rate/mass loss rate) of 0.6, 3 and 6 \\%, respectively. Walder \\& Folini (2000) also presented a nice review of wind dynamics in symbiotics binaries, with an emphasis on the influence of the radiation field of the accretor. More recently, Dumm et al. (\\cite{Dumm00}) presented three-dimensional Eulerian isothermal simulations in order to represent the symbiotic system RW Hya which seems to present an accretion wake, as indicated by the reduced UV flux observed at phase $\\Phi = 0.78$. They found that 6\\% of the M-giant wind is captured by the companion. In this paper we perform three-dimensional numerical simulations of mass transfer by wind in a binary system with a mass ratio of unity. The full gravitational forces of the two components are taken into account. This is therefore different from the e.g. TBJ96 or Dumm et al. (2000) studies, where in order to account for the not very well known acceleration mechanism of the wind, the gravitational force of the primary is partially or totally reduced. Here, we will concentrate on the derivation of the mass accretion ratio. We will leave for a following paper the physical discussion of the flow, including a discussion of angular momentum loss from the system. ", "conclusions": "\\begin{enumerate} \\item The flow pattern is classified by means of the normal speed of gas on the critical Roche surface, $V_R$. For $V_R<0.4A\\Omega$, we have a RLOF type flow, while for $V_R>0.7A\\Omega$, wind accretion flows are realized. In the intermediate parameter range, $0.4A\\Omega10^{12}$\\,cm$^{-3}$). Densities this high are more easily explained in terms of compact X-ray emitting regions (e.g. Sanz-Forcada et al. 2003). Brickhouse et al. (2001) find evidence for compact, quiescent coronae in active stars in a {\\em Chandra}/HETG study of the contact binary, 44i~Boo (G0V~\\&~G0V), spanning over 2.5\\,${P}_{\\rm rot}$. Flaring emission is filtered out and velocity shifts in line centroids of emission line profiles are measured as a function of rotation phase. Rotational modulation in the quiescent corona is detected, indicating a compact emitting area near the pole of the primary. We want to follow up on this latter study in order to learn whether the picture of compact stellar coronae is applicable to other active cool stars or whether it is exclusive to contact binaries. \\subsection{AB Doradus} AB Doradus (K0V, $m_V$=6.9, ${P}_{\\rm rot}$=0.51\\,d, v$_e \\sin i$=90\\,km\\,s$^{-1}$) is an active, rapidly rotating single star that has recently arrived onto the main sequence. It is an ideal candidate for spectroscopic mapping techniques such as Doppler imaging as its 12.4\\,hour rotation period means that a large fraction of its surface can be tracked in one night, and its rotationally broadened spectral line profiles enable detailed surface maps to be reconstructed ($\\sim 3$\\deg\\ longitude resolution at the equator). Spot maps of AB Dor's surface typically show a dark spot covering its pole (polar cap) with smaller spots co-existing at lower latitudes (e.g. Donati et al. 1999). AB Dor is also bright in X-ray wavelengths, $L_{x}/L_{bol}~10^{-3}$, and shows strong signs of coronal variability. ROSAT X-ray light curves of AB Dor indicate rotational modulation at the 5-13\\% level (Kuerster et al. 1997). ", "conclusions": "Fig.~3 shows the X-ray light curve for AB Dor. We find that the light curve varies by 30\\%. Initial periodogram analyses indicate that a fraction of this is indeed due to rotational modulation. We will use the method used by K\\\"urster et al. (1997) to filter this X-ray light curve in order to distinguish the rotationally modulated component. This rotationally modulated component of the light curve will be compared with the predicted light curve in Fig.~1. Our 3-D coronal model of AB Dor can also be refined, we can compensate for missing magnetic field information (e.g. in the polar cap). However, only a unipolar spot should significantly modify the coronal field distribution (McIvor et al. 2003). We will also evaluate the effect on the predicted X-ray lightcurve if the global magnetic field is non-potential using the technique described by Hussain et al. (2002). Measurements of the centroids of the best resolved spectral lines (the summed profile derived using the above method and O\\,{\\sc viii}\\,18\\,\\AA\\ line) show evidence of rotational modulation that repeats from one rotation cycle to the next. Even though both the O\\,{\\sc viii}\\,18\\,\\AA\\ and the summed line profile show the same trend, the uncertainty associated with our velocity shift measurements makes this result somewhat inconclusive. This pattern of rotational modulation would indicate that the bulk of quiescent X-ray emission in AB Dor (at this epoch) originated in localised regions, most likely at high latitudes ($\\sim$60$^{\\circ}$). Emission originating at the pole would show no rotational modulation. We will add more lines from the longer wavelengths ($>88$\\,\\AA) to reduce the error bars on this measurement." }, "0403/astro-ph0403486_arXiv.txt": { "abstract": "We present a spectroscopic analysis of nearly 8000 late-type dwarfs in the Sloan Digital Sky Survey. Using the H$\\alpha$ emission line as an activity indicator, we investigate the fraction of active stars as a function of spectral type and find a peak near type M8, confirming previous results. In contrast to past findings, we find that not all M7-M8 stars are active. We show that this may be a selection effect of the distance distributions of previous samples, as the active stars appear to be concentrated near the Galactic Plane. We also examine the activity strength (ratio of the luminosity emitted in H$\\alpha$ to the bolometric luminosity) for each star, and find that the mean activity strength is constant over the range M0-M5 and declines at later types. The decline begins at a slightly earlier spectral type than previously found. We explore the effect that activity has on the broadband photometric colors and find no significant differences between active and inactive stars. We also carry out a search for subdwarfs using spectroscopic metallicity indicators, and find 60 subdwarf candidates. Several of these candidates are near the extreme subdwarf boundary. The spectroscopic subdwarf candidates are redder by $\\sim 0.2$ magnitudes in $g-r$ compared to disk dwarfs at the same $r-i$ color. ", "introduction": "Low mass stars with late spectral types (M,L) are the majority constituent of the Galaxy by number. Their main-sequence lifetimes are much greater than the current age of the Universe and they therefore serve as useful probes of Galactic star formation history in the local solar neighborhood (Gizis, Reid \\& Hawley 2002). They also encompass many important regions of stellar parameter space, including the onset of complete convection in the stellar interior, the onset of significant electron degeneracy in the core, and the formation of dust and subsequent depletion of metals onto dust grains in the stellar atmosphere. Of particular interest is the fact that many late-type stars have strong surface magnetic fields (Johns-Krull \\& Valenti 1996) that heat the outer atmosphere above the photosphere, and lead to observable emission from the chromosphere (e.g. Ca II and H Balmer series lines), the transition region (e.g resonance lines of abundant ions such as C IV), and the corona (e.g. thermal soft X-rays). The physics that controls the production of magnetic fields in low mass stars is not well understood, as the lack of a radiative-convective boundary layer precludes storing large scale fields as in a solar-type dynamo. However, recent work on turbulent dynamo mechanisms (Bercik et al. 2004) may soon provide new insight into the formation and properties of magnetic fields in low mass stars. The present study seeks to place strong empirical constraints on the magnetic activity in these stars, by measuring optical H$\\alpha$ emission in a sample of nearly 8000 low mass dwarfs. Our results will aid in the interpretation of the models and provide a connection to the physical processes occurring in the atmospheres of the stars. The H$\\alpha$ emission line is produced by collisional excitation in the relatively dense chromospheres of these low mass, high gravity stars. Located in the red region of the optical spectrum, it is the strongest and best-studied indicator of magnetic activity in late type stars (in contrast to solar type stars which are usually studied in the Ca II H and K resonance lines). Previous results indicate that the fraction of M dwarfs with H$\\alpha$ emission increases from early to mid M spectral types (Joy \\& Abt 1974; Hawley, Gizis \\& Reid 1996), reaches a peak near type M7 where essentially all stars are active, and declines toward later types (Gizis et al. 2000a). The increasing fraction in the early to mid M types may be an age effect such that the activity lasts longer in the mid M types (Hawley, Reid \\& Tourtellot 2000a; Gizis et al. 2002), while the decline toward late M types may reflect difficulty in producing and/or maintaining surface magnetic fields due to the physics of turbulent dynamos and/or the increasingly neutral atmosphere (Hawley, Reid \\& Gizis 2000b; Fleming, Giampapa \\& Schmitt 2000; Mohanty et al. 2002). It does not appear that the atmosphere changes character such that H$\\alpha$ is no longer produced when magnetic activity is present in the later type stars. Indeed the ratio of H$\\alpha$ to transition region (C IV) and soft X-ray emission appears to be similar throughout the M dwarf spectral sequence (Hawley \\& Johns-Krull 2003). Previous work also shows that the activity strength, measured by the ratio of the luminosity in H$\\alpha$ to the bolometric luminosity (L$_{\\rm{H}_\\alpha}$/L$_{bol}$), is nearly constant (with large scatter) through the M0-M6 range (Hawley et al. 1996) and declines at later types (Burgasser et al. 2002, Cruz \\& Reid 2002). However, the sample of stars on which these conclusions are based numbers less than 100 for types M7 and later, due to the difficulty of obtaining spectra for these faint objects. Our sample in this crucial spectral type range is an order of magnitude larger, comprising more than 1000 stars with types between M7 and L0, and nearly 8000 stars between M0 and L0. This uniformly acquired and reduced sample from the Sloan Digital Sky Survey (SDSS), with well-characterized uncertainties, now allows us to obtain statistically significant magnetic activity results for the entire M dwarf spectral type range. In addition, we compare the photometric colors of the active and inactive stars to investigate whether the active stars are typically bluer at the same spectral type, an effect attributed to an increase in plage areas on active stars by Amado \\& Byrne (1997). Alternatively, such a correlation might indicate that micro-flaring activity plays a significant role in heating the outer atmospheres of active M dwarfs (cf. G\\\"{u}del et al. 2003). Our SDSS spectral sample is also well-suited for investigation of the incidence of low mass subdwarfs in the local neighborhood. Gizis (1997) showed that M subdwarfs could be identified by comparing the metallicity and gravity sensitive molecular bands of CaH and TiO. Using his technique, we identify dozens of new M subdwarf and extreme subdwarf candidates, and examine their colors in the SDSS $ugriz$ filters. ", "conclusions": "Using a sample of more than 22000 candidates from the SDSS spectroscopic database, 7840 late type stars were identified and used to investigate the magnetic activity in M and early L dwarfs. This sample is much larger than all previous samples used for this purpose. Our results show that: \\begin{enumerate} \\item{The fraction of active stars rises monotonically from spectral type M0 to M8 and then declines monotonically to the latest types we measured (L3-L4).} \\item{Only $\\sim 70\\%$ of the M7 stars in our sample are active, which differs from previous studies showing 100\\% activity at spectral type M7. We suggest that the answer to this discrepancy may lie in the distance distributions of the samples used. Our results indicate the possibility that stars near the Galactic Plane are more likely to be active, as expected if activity depends on age.} \\item{The activity strength, as measured by ${\\rm L}_{\\rm{H}\\alpha}/{\\rm L}_{bol}$, is approximately constant from types M0-M5, and declines at later types. The width of the distribution is relatively narrow at early types but broadens considerably at type M6 and later.} \\item{The average $(u-g)$ color is slightly bluer for active stars compared to inactive ones, while the average $(g-r)$ color is slightly redder. However, these differences are not statistically significant.} \\item{60 new subdwarf candidates are identified spectroscopically by comparing the strengths of the TiO and CaH molecular bands. Several of the candidates lie near the extreme subdwarf boundary and may be interesting targets for future observations.} \\item{The subdwarf candidates have similar colors to the disk dwarfs except in the $(g-r)$ color, where the median subdwarf $(g-r)$ is almost 0.2 magnitudes redder than disk dwarfs at the same $(r-i)$ color. This effect allows subdwarfs to be chosen photometrically from the SDSS database.} \\end{enumerate}" }, "0403/astro-ph0403165_arXiv.txt": { "abstract": "{{ We present a combined radio, optical, and X-ray study of the nearby LINER galaxy NGC\\,1052. Data from a short (2.3\\,ksec) {\\it CHANDRA} observation of NGC\\,1052 reveal the presence of various jet-related X-ray emitting regions, a bright compact core and unresolved knots in the jet structure as well as an extended emitting region inside the galaxy well aligned with the radio synchrotron jet-emission. The spectrum of the extended X-ray emission can best be fitted with a thermal model with $kT = (0.4-0.5)$\\,keV, while the compact core exhibits a very flat spectrum, best approximated by an absorbed power-law with $N_{\\rm H} = (0.6-0.8) \\times 10^{22}\\,{\\rm cm^{-2}}$. We compare the radio structure to an optical ``structure map'' from a {\\it Hubble Space Telescope} ({\\it HST}) observation and find a good positional correlation between the radio jet and the optical emission cone. Bright, compact knots in the jet structure are visible in all three frequency bands whose spectrum is inconsistent with synchrotron emission.} ", "introduction": "NGC\\,1052 is a nearby\\footnote{$D = $ 22.6\\,Mpc (assuming $z =$ 0.0049 (Knapp et al. \\cite{Kna78}) and $H_0 = 65$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$). At this distance 1\\,arcsec corresponds to $\\sim$110\\,pc.} elliptical galaxy which harbors a low-luminosity active galactic nucleus (LLAGN) in its very center (L$_{1-100{\\rm GHz}}=4.4 \\times 10^{40}$\\,erg\\,s$^{-1}$; Wrobel \\cite{Wro84}). { It hosts} a two-sided radio jet emanating from the nucleus and reaching out to { kiloparsec-scales} which is, however, still fully enclosed within the stellar body of the optical galaxy. In the optical, the spectrum of NGC\\,1052 is characterized by strong forbidden lines from low-ionization states which has made NGC\\,1052 the prototypical LINER (low-ionization nuclear emission line region; Heckman \\cite{Hec80}) galaxy. As for LINERs in general, it has long been argued whether these low-ionization lines in NGC\\,1052 are excited by a central photo-ionizing source (e.g., Gabel et al. \\cite{Gab00}) or if shock heating is the dominant mechanism (e.g., Sugai \\& Malkan \\cite{Sug00}). While there is overwhelming evidence for the presence of an active galactic nucleus (AGN) in NGC\\,1052, the role of shocks in this galaxy is still unclear. The improved angular resolution in X-rays offered by {\\it CHANDRA} makes it possible for the first time to image the distribution of X-ray emission on the same scales as accomplished by connected radio interferometers, e.g., MERLIN (Multi-Element Radio-Linked Interferometer Network). Disentangling the contributions of compact (i.e., $<$1\\,arcsec) nuclear and extended (i.e., $>$1\\,arcsec) X-ray emitting regions to the total amount of X-ray emission of NGC\\,1052 can serve as an important tool to study the interaction between the radio jet plasma and the ambient interstellar medium. NGC\\,1052 has been observed by all major X-ray missions of the pre--{\\it CHANDRA} era, like {\\it Einstein} (Mc Dowell \\cite{McD94}), {\\it ASCA} and {\\it ROSAT} (Weaver et al. \\cite{Wea99}), and {\\it Beppo Sax} (Guainazzi \\& Antonelli \\cite{Gua99}). For these X-ray missions, NGC\\,1052 appeared as a point-like X-ray source. { The X-ray spectrum of NGC\\,1052 is extremely flat, a finding that led to the proposal of an advection dominated accretion flow (ADAF) as the origin of the observed X-ray emission (Guainazzi et al. \\cite{Gua00}).} To model the AGN X-ray spectrum above $E \\simeq $ 2\\,keV absorbing column densities in excess of $10^{23}$\\,cm$^{-2}$ have been discussed, supporting the idea of a high density obscuring torus, comparable to column densities found in other AGN (e.g., Malizia et al. \\cite{mali97}, Risaliti et al. \\cite{risa02}). Independent evidence for the existence of an obscuring torus at the center of NGC\\,1052 is obtained from Very Long Baseline Interferometry (VLBI) observations in the radio regime: on { parsec-scales} NGC\\,1052 exhibits a twin jet structure with a prominent emission gap between both jets (see e.g., Kadler et al. \\cite{Kad03b}). The inner part of the western jet shows a strongly inverted radio spectrum, which was first discovered by Kellermann et al. (\\cite{Kel99}) (see also Kameno et al. \\cite{Kam01}). The cm-wavelength spectral index in this central region is larger than 2.5, exceeding the theoretical limit for synchrotron self-absorption. Combined studies of the core region of NGC\\,1052 in the radio and X-ray regime are of essential importance { for constraining the physical properties of the { parsec-scale} radio jet and the obscuring torus as well as to determine the nature of the nuclear X-ray emission.} In this paper we present a combined radio, optical, and X-ray study of the jet-related emission in NGC\\,1052 on arcsecond scales. In particular, we focus on the soft X-ray excess in the source-spectrum below $E = $ 2\\,keV. This soft component was identified first by Weaver et al. (\\cite{Wea99}) based on {\\it ROSAT} PSPC data. {\\it CHANDRAs} superior angular resolution makes it possible to present evidence that this soft excess emission is associated with the well known radio jet. In Sect.~\\ref{obs} we present the {\\it CHANDRA}, MERLIN, and {\\it HST} data as well as their reduction. In Sect.~\\ref{multi-structure} we discuss the arcsecond-scale morphology of NGC\\,1052 in the radio, optical, and X-ray regime and the correlations between the different wave bands. In Sect.~\\ref{spectroscopy} we derive models for the nuclear and extended X-ray emission and Sect.~\\ref{summ} summarizes our conclusions. ", "conclusions": "\\label{summ} The {\\it CHANDRA} data provide for the first time direct evidence for { jet-associated X-ray emission in NGC\\,1052.} The diffuse, extended X-ray emission can be best approximated with a thermal plasma model with { $kT \\sim 0.4-0.5$\\,keV}. This temperature is consistent with the thermal component found earlier by Weaver et al. (1999) using {\\it ASCA} and {\\it ROSAT} data. Its absorbed flux is only { $\\sim 3$\\%} of the nuclear X-ray emission { but the intrinsic (absorption corrected) extended emission might contribute up to 14\\% to the total unabsorbed X-ray flux of NGC\\,1052.} Because of the { considerable} pile-up degradation of the {\\it CHANDRA} data, no firm conclusions on the photon index of the nucleus spectrum can be deduced. The derived column density of hydrogen towards the compact X-ray core (depending on the applied model) { of $0.5-0.8 \\times 10^{22}$\\,cm$^{-2}$ is in good agreement with the absorbing column density of ionized material towards the VLBI-jet derived by Kadler et al. (\\cite{Kad02}) and Kadler et al. (in prep.). This suggests that the nuclear X-ray emission of NGC\\,1052 might be produced on the same scales as the parsec-scale structures imaged by VLBI at high frequencies.} The detection of a diffuse region of X-ray emitting gas with a thermal spectrum and the same extent as the kiloparsec-scale radio jet suggests that jet-triggered shocks might play an important role in NGC\\,1052. In such a model the kinetic power of the radio jet is partially converted into X-ray emission. The optical morphology in the H$\\alpha$ filter substantiates this picture as was noted earlier by Allen et al. (\\cite{All99}). The alignment of the radio jet and the optical emission cone visible in Fig.~\\ref{fig:structure_map} implies that the ionization cone might be drilled out by the radio jets, resulting in a predominantly shock-excited, conical narrow-line region (see e.g., Dopita \\cite{Dop02}). Shocks might occur also on larger scales giving rise to the soft thermal X-ray emission associated with the radio jet/lobe structure in NGC\\,1052. A rough estimate (see Kadler et al. \\cite{Kad03a}) shows that the soft thermal X-ray spectrum associated with the radio jet of NGC\\,1052 can be explained in terms of the kinetic jet power being partially converted into X-ray emission originating in shocks driven into the ambient medium. (A more detailed model of the relation between jet-driven shock-activity and the spectral shape of the extended X-ray emission in NGC\\,1052 will be discussed in a forthcoming paper.) The comparison of the large-scale distribution of radio emission in NGC\\,1052 between two epochs separated by $\\sim$ 15 years indeed shows activity on { kiloparsec-scales}. This substantiates the idea that shocks in the interstellar medium form at the working surfaces of active regions (hotspots and knots). Moreover, recent numerical simulations (e.g., Zanni et al. \\cite{Zan03}) show that jets in radio galaxies can inflate over-pressured cocoons that drive shocks into the ambient gas resulting in morphologies (in the case of weak shocks) very similar to what is observed in NGC\\,1052: a cavity of hot X-ray emitting gas in conjunction with a local deficit of X-ray emission around the hotspots. A deeper {\\it CHANDRA} observation with an improved photon statistic { compared to} the observation discussed here { would} provide both a higher sensitivity to the weak diffuse emission and a higher resolution. Additionally, the full resolution of {\\it CHANDRA} of $\\sim 0.5$\\,arcsec would allow one to study in more detail the connection between the knots in the diffuse X-ray emission and the optical emission knots." }, "0403/astro-ph0403353_arXiv.txt": { "abstract": "The large-angle (low-$\\ell$) correlations of the Cosmic Microwave Background exhibit several statistically significant anomalies compared to the standard inflationary cosmology. We show that the quadrupole plane and the three octopole planes are far more aligned than previously thought ($99.9$\\% C.L.). Three of these planes are orthogonal to the ecliptic at $99.1$\\% C.L., and the normals to these planes are aligned at $99.6$\\% C.L.~with the direction of the cosmological dipole and with the equinoxes. The remaining octopole plane is orthogonal to the supergalactic plane at $99.6$\\% C.L. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403679_arXiv.txt": { "abstract": "s{We are undertaking the first systematic study of the prevalence of AGN activity in a large sample of high redshift galaxy clusters. Local clusters contain mainly red elliptical galaxies, and have little or no luminous AGN activity. However, recent studies of some moderate to high redshift clusters have revealed significant numbers of luminous AGN within the cluster. This effect may parallel the Butcher-Oemler effect -- the increase in the fraction of blue galaxies in distant clusters compared to local clusters. Our aim is to verify and quantify recent evidence that AGN activity in dense environments increases with redshift, and to evaluate the significance of this effect. As cluster AGN are far less prevalent than field sources, a large sample of over 120 cluster fields at $z > 0.1$ has been selected from the Chandra archives, and is being analysed for excess point sources. The size of the excess, the radial distribution and flux of the sources and the dependence of these on cluster redshift and luminosity will reveal important information about the triggering and fueling of AGN. } ", "introduction": " ", "conclusions": "" }, "0403/nucl-th0403066_arXiv.txt": { "abstract": "We use Landau's theory of a normal Fermi liquid to derive expressions for the static response of a system with a general tensor interaction that conserves the total spin and the total angular momentum of the quasiparticle-quasihole pair. The magnetic susceptibility is calculated in detail, with the inclusion of the center of mass tensor and cross vector terms in addition to the exchange tensor one. We also introduce a new parametrization of the tensor Landau parameters which significantly reduces the importance of high angular harmonic contributions. For nuclear matter and neutron matter we find that the two most important effects of the tensor interaction are to give a contribution from multipair states and to renormalize the magnetic moments. Response to a weak probe may be calculated using similar methods, replacing the magnetic moments with the matrix elements of the weak charges. ", "introduction": "For systems with central interactions, Landau's theory of normal Fermi liquids provides an economical way of characterizing many low-temperature properties. The theory was applied to atomic nuclei by Migdal and collaborators, and in that work it was generally assumed that the non-central contributions to the effective nucleon-nucleon interaction were small \\cite{migdal}. The generalization of Landau theory to include effects of the tensor force was made by D\\c{a}browski and Haensel \\cite{dabrowski2,dabrowski1,haensel2}. Subsequently, estimates of tensor contributions to the effective interaction were made in Refs.\\ \\cite{baeckman,schwenk,schwenk2,dickhoff}. The stimulus for the present work arose in the context of astrophysics. In the physics of collapse and the subsequent evolution of a neutron star, the properties of neutrinos in dense matter are a key ingredient \\cite{janka}. As demonstrated in Refs.\\ \\cite{sawyer,iwamoto,raffelt1,raffelt2,raffelt3,burrows,prakash,prakashreddy}, neutrino scattering and absorption rates are sensitive to nucleon-nucleon interactions, especially their spin-dependent parts. Direct calculation of effective interactions is difficult, but for systems with only central interactions, information about the effective interaction at long wavelengths may be obtained directly from a knowledge of static properties of the matter. For example, the spin susceptibility of a single-component Fermi liquid with two spin states is given by \\beq \\chi=\\frac{\\mu_0^2 N(0)}{1+G_0}, \\label{Landauchi} \\eeq where $\\mu_0$ is the magnetic moment of a particle in free space, $N(0)=m^* p_{\\rm F}/\\pi^2\\hbar^3$ is the density of states per unit volume at the Fermi surface, $m^*$ is the quasiparticle effective mass, $p_{\\rm F}$ is the Fermi momentum and $G_0$ is the Landau parameter describing the isotropic part of the spin dependent contribution to the quasiparticle interaction. Since calculations of the magnetic susceptibility of neutron matter exist \\cite{Fantoni}, it is relevant to ask to what extent it is possible to deduce properties of the effective interaction from such data. A related question is the extent to which tensor contributions to quasiparticle energies and interactions alter neutrino scattering rates, which were previously calculated neglecting tensor effects in Ref. \\cite{iwamoto}. As shown in earlier work \\cite{olsson}, the tensor force influences the static response of a system in a number of different ways. One is that the magnetic moment of a quasiparticle is different from its value for a particle in vacuum. In particular, the magnetic moment is not a scalar, as it is for systems with central forces only. As long ago as 1951, Miyazawa \\cite{miyazawa} calculated explicitly how the magnetic moment of a nucleon would be modified by the tensor interaction due to one pion exchange and more recent discussions may be found in e.g. the review \\cite{arima} and Ref. \\cite{riska}. A second effect is that the quasiparticle interaction contains explicit tensor contributions. Following Refs. \\cite{dabrowski1,dabrowski2,haensel2}, it has generally been assumed that these have an exchange-tensor structure similar to that of the one-pion exchange interaction. However, Schwenk and Friman \\cite{schwenk2} have pointed out recently that the one-pion exchange interaction, when acting in second order, can give rise to contributions to the effective interaction which have a different structure. In their paper they evaluated these induced interaction contributions to the quasiparticle interaction and found that the exchange-tensor term in the quasiparticle interaction is much reduced, and that terms of a different structure can be of a comparable magnitude to the exchange-tensor ones. A third effect of the tensor interaction is that there are multipair contributions to the magnetic susceptibility. In this paper we begin by deriving a general expression for the magnetic susceptibility in Sec.\\ \\ref{Sec:response}, taking into account tensor contributions to the magnetic moment, and tensor contributions to the effective interaction which are completely general for an interaction which conserves the total angular momentum and the total spin of the quasiparticle-quasihole pair. This represents a generalization of the earlier calculation of the magnetic susceptibility by Haensel and Jerzak \\cite{haensel}, who took into account the effect of the exchange-tensor contribution to the quasiparticle interaction. Another issue that we address is how to parametrize tensor contributions to the quasiparticle interaction (Sec.\\ \\ref{Sec:tensor}). The scheme usually employed in the past suffers from the disadvantage that it is generally necessary to take into account high angular harmonic contributions to the interaction. We present an alternative parametrization for which higher harmonic terms play little role. In Sec.\\ \\ref{Sec:ME} we evaluate the matrix elements needed for the calculation of the susceptibility and estimate the magnitudes of the different contributions. In Sec.\\ \\ref{Sec:SNM} we extend the result to symmetric nuclear matter and to responses with other spin and isospin properties, for example that to weak interactions. Sec.\\ \\ref{Sec:conclusions} contains concluding remarks. ", "conclusions": "\\label{Sec:conclusions} In this paper we have derived an expression for the magnetic susceptibility of a Fermi liquid for a general interaction which conserves the total angular momentum and the total spin of the quasiparticle-quasihole pair. Apart from the contribution from single quasiparticle-quasihole pairs, which may be calculated using Landau theory, the susceptibility also contains a contributions from excitation of multipair states. In addition to the exchange tensor term usually included, we have taken into account the new tensor terms found in Ref. \\cite{schwenk2}. We have also introduced an alternative parametrization of the exchange tensor interaction which has the advantage of reducing the importance of high angular harmonic contributions. For neutron matter we find that the most important effects of the tensor force are to renormalize the magnetic moment and to give a contribution from multipair states. For symmetric nuclear matter one can draw the same conclusion, the bound on the multipair contribution is, however, much smaller, but it is not clear how close the actual contribution to the susceptibility, from multipairs, is to this lower bound. The formalism may easily be extended to calculate the response to weak probes: the only difference being that the magnetic moments must be replaced by the corresponding matrix elements of the weak charges." }, "0403/astro-ph0403308_arXiv.txt": { "abstract": "We build template maps for the polarized Galactic--synchrotron emission on large angular scales (FWHM~=~7$^\\circ$), in the 20-90~GHz microwave range, by using WMAP data. The method, presented in a recent work, requires a synchrotron total intensity survey and the {\\it polarization horizon} to model the polarized intensity and a starlight polarization map to model polarization angles. The basic template is obtained directly at 23~GHz with about 94\\% sky--coverage by using the synchrotron map released by the WMAP team. Extrapolations to 32, 60 and 90~GHz are performed by computing a synchrotron spectral index map, which strongly reduces previous uncertainties in passing from low (1.4~GHz) to microwave frequencies. Differing from low frequency data, none of our templates presents relevant structures out of the Galactic Plane. Our map at 90~GHz suggests that the synchrotron emission at high Galactic latitudes is low enough to allow a robust detection of the $E$--mode component of the cosmological signal on large--scale, even in models with low--reionization ($\\tau = 0.05$). Detection of the weaker $B$--mode on the largest scales ($\\ell < 10$) might be jeopardized unless the value $\\tau = 0.17$ found by WMAP is confirmed, and $T/S > 0.1$. For lower levels of the gravitational--wave background the $B$--mode seems to be accessible only at the $\\ell \\sim 100$ peak and in selected low--synchrotron emission areas. ", "introduction": "\\label{intro} Recent results by DASI (Kovac et al. 2002) and WMAP (Bennett et al 2003a) have opened a new window in studying both the Cosmic Microwave Background (CMB) and the Galaxy. The polarized signal detected with the DASI experiment, as well as the WMAP measurement of the temperature--polarization cross--spectrum $C^{TE}_\\ell$, emphasized the importance of studying CMB Polarization (CMBP). In particular, the discovery of a relevant signal on large angular scales performed by WMAP has provided evidence of an unexpectedly early reionization era (Kogut et al. 2003). Furthermore, WMAP provided the first all--sky total--intensity maps of the microwave Galactic emission, allowing an insight into synchrotron, dust and free-free components of the Galaxy (Bennett et al. 2003b). A sound measurement of CMBP requires good knowledge of polarized foregrounds (both Galactic and extragalactic), which are potentially more dangerous than in total intensity. This is even more true for the $B$--mode, retaining information on the gravitational--wave background (Kamionkowski \\& Kosowski, 1998), whose emission level is expected to be orders of magnitude below that of the $E$-mode. Among all foreground components, the Galactic synchrotron radiation is expected to be the most relevant up to 100~GHz. The lack of data in both the CMB frequency range and at high Galactic latitudes makes the building of templates necessary (Kogut \\& Hinshaw 2000, Giardino et al. 2002, Bernardi et al. 2003a, hereafter B03). These can substantially help in studying how to extract CMB maps from contaminated signals (e.g. see Bennett et al. 2003b). In B03 we presented a method modelling the Galactic polarized--synchrotron emission by using the radio--continuum total intensity surveys available at low frequency ($< 1.4$~GHz) to model the polarized intensity, and starlight polarization optical data as a template for polarization angles. Template maps obtained with this method are virtually free of Faraday rotation. However, when extrapolating low frequency data (the only available before the latest WMAP release) to the cosmological window, uncertainties arise on the spectral index to be used. In B03 we used the mean spectral index derived by Platania et al. (1998) up to 19~GHz, since no information were available at higher frequencies. In addition, variations across the sky were too poorly known to be taken into account. The WMAP release of total--intensity synchrotron maps in the cosmological window gives us the possibility to significatively improve our templates. Actually, the application of our method directly at frequencies of interest for CMBP fully avoids extrapolation uncertainties. Our templates are built with FWHM~$=7^{\\circ}$ at the three frequencies of 23, 32 and 90~GHz interesting for the SPOrt experiment (Cortiglioni et al. 2004), one of whose aims is studying the CMBP foregrounds. We also generate a template at 60~GHz, this frequency being of interest for other CMB experiments (WMAP, Planck). All of the templates cover large fractions of the sky, namely 94\\% at 23~GHz and 92\\% at 32, 60 and 90~GHz. In addition, we calculate polarized--syncrotron angular power spectra both for the full-sky templates and for the high Galactic latitude emission, discussing possible effects on the measurement of CMBP $E$ and $B$--mode spectra. The paper is organized as follows: in Section~\\ref{model} we apply our method to WMAP data at 23~GHz and provide a Galactic polarized synchrotron template at the same frequency; in Section~\\ref{extr} we extrapolate our results up to 90~GHz; in Section~\\ref{tot_aps} we describe the angular power spectra of our full--sky templates at the different frequencies, in Section~\\ref{for_aps} we compute the angular power spectra for the high Galactic latitude emission and compare synchrotron to CMBP and, finally, conclusions are presented in Section~\\ref{conc}. ", "conclusions": "\\label{conc} In this work we generate template maps of polarized Galactic synchrotron emission at 23, 32, 60 and 90~GHz using WMAP data. We apply the method developed in B03 to the 23~GHz total intensity synchrotron map released by the WMAP team. This greatly improves the B03 templates avoiding the uncertainties in the extrapolation from low frequency. As for the previous work, the angular resolution is limited to FWHM=$7^\\circ$ mainly due to the starlight polarization data sampling. The basic template (23--GHz--W) is constructed at 23~GHz applying the method directly to the WMAP data and it covers almost all the sky (94\\%). Differing from the 1.4~GHz B03 template where the Northern Galactic Spur represents an important feature, 23--GHz--W does not present relevant structures out of the Galactic Plane. This replicates what already occurs in total intensity and leaves most part of the sky with low synchrotron emission and useful for CMBP investigations. To achieve the templates at the other frequencies a spectral index map is required to extrapolate the 23--GHz--W result. Differing from Bennett et al. (2003b), which derived a spectral index map from 0.408 to 23~GHz with no free--free subtraction in the 0.408~GHz map, we computed the frequency behaviour using {\\it pure} synchrotron maps at 1.4 and 23~GHz. The 23~GHz one is that released by the WMAP team. The 1.4~GHz map has been achieved from low frequency surveys by applying a Dodelson-like component separation. We use the Haslam et al. 0.408~GHz and the Reich 1.4~GHz surveys for the Northern sky and the Haslam et al. and the Jonas et al. 2.3~GHz ones for the Southern Hemisphere. The results is the first almost all-sky spectral index map of the Galactic synchrotron emission in the 1.4--23~GHz range. Furthermore, we add an extra 0.5 to the indeces to account for the steepening found by Bennett et al. (2003b) in the 23--41~GHz range and this spectral index map is used to compute template maps at 32, 60 and 90~GHz covering 92\\% of the sky. The spectral index map provides steeper indexes at high Galactic latitudes so that the resulting templates have the Galactic Plane emission more and more dominant from 23~GHz up to 90~GHz. This difference also explains the disappearing of the Northern Galactic Spur in our templates. Unclear, instead, it is the reason for such a flat frequency spectrum found in the Galactic Plane, it being either intrinsic to the synchrotron emission (for instance due to a young population of relativistic electrons) or due to a residual thermal component. Beyond the lack of relevant structures at high Galactic latitudes, the differences among the templates of this work and the B03 one also concern the APS. APS of our full--sky templates are dominated by the emission from the Galactic Plane and flatter slopes, namely in the 1.4--1.6 range for $C^E_\\ell$, 0.5--0.8 for $C^B_\\ell$ and 1.2--1.5 for $C^P_\\ell$. This general flattening is still not clear. A further unclear point is that the synchrotron APS slopes (in particular $C^B_\\ell$ and $C^P_\\ell$) seem to flatten with increasing frequency. If confirmed, these two aspects would confirm that low frequency APS cannot be straightforwardly extrapolated from low to microwave frequencies. The situation is less complex at high Galactic latitudes. In fact, limiting the analysis out of the Galactic Plane, a steepening of the slopes occurs, which become more similar to the B03 ones, and no significative dependency from the frequency is found. Interesting conclusions are drawn by the comparison of the high Galactic latitude portion of our templates to CMBP as from the {\\it concordance model} of the WMAP first--year data. Indeed, the CMBP $E$--mode spectrum is about 2 orders of magnitude above the synchrotron signal in our template at 90~GHz. Synchrotron is thus unlikely to contaminate the cosmological signal. Furthermore, even a less re-ionized Universe ($\\tau = 0.05$) should be dominant over the Galactic emission. The situation looks different for the CMBP $B$--mode. Our template predicts that this cosmological signal might be accessible at largest scales ($\\ell < 10$) only for a Tensor to Scalar fluctuation power ratio $T/S > 0.1$. To investigate models with lower values of $T/S$ one should restrict to selected low synchrotron emission areas large enough to detect the $\\ell \\sim 100$ peak. An example is the sky patch observed by the BOOMERanG experiment where models with $T/S > 0.01$ seem to be accessible (Bernardi et al. 2003b). Our templates obviously rely on the assumption that WMAP's dust-correlated signal at 22~GHz is genuine synchrotron, as supported by Bennet et al. (2003b). However several authors (de Oliveira--Costa et al. 2003, Banday et al. 2003, Lagache 2003) recently support the opposite view, that WMAP's data may be more consistent with anomalous dust emission. For instance, while the spinning-dust model originally proposed by Draine \\& Lazarian (1998) may meet some difficulty, emission by very small grains is supported by Lagache (2003). Considering that dust polarization level is likely less than 5\\% rather than the 15-30\\% of the synchrotron, a measurement of polarization at 22-32~GHz could help to decide on this issue: should the polarization level be found a factor $\\sim 5$ below the predictions of our template, the role of anomalous dust emission would be supported. Also, in such a case, the contamination of CMBP maps at 90~GHz would be even smaller than considered in our discussion above, and prospects for the detection of a cosmological B mode would be slighly better." }, "0403/astro-ph0403414_arXiv.txt": { "abstract": "We report the search for intracluster light in four Abell Type II/III (non-cD) galaxy clusters: Abell 801, 1234, 1553, \\& 1914. We find on average that these clusters contain $\\sim$ 10\\% of their detected stellar luminosity in a diffuse component. We show that for two of the clusters the intracluster light closely follows the galaxy distribution, but in the other two cases, there are noticeable differences between the spatial distribution of the galaxies and the intracluster light. We report the results of a search for intracluster tidal debris in each cluster, and note that Abell~1914 in particular has a number of strong tidal features likely due to its status as a recent cluster merger. One of the Abell~1914 features appears to be spatially coincident with an extension seen in weak lensing maps, implying the feature traces a large amount of mass. We compare these results to numerical simulations of hierarchically-formed galaxy clusters, and find good general agreement between the observed and simulated images, although we also find that our observations sample only the brightest features of the intracluster light. Together, these results suggest that intracluster light can be a valuable tool in determining the evolutionary state of galaxy clusters. ", "introduction": "As the most massive gravitationally bound structures in the universe, galaxy clusters stand to teach us much about the hierarchical assembly of matter in the universe. Clusters exhibit a wide variety of structural properties, from massive, X-ray luminous clusters dominated by early type galaxies (i.e., Coma) to irregular, spiral-rich clusters like the Ursa Major cluster, down to poor clusters and loose groups. The fact that many clusters are still obviously in the process of assembly can be seen via many tracers of substructure, such as X-ray isophotes, gravitational lensing maps, and kinematic and spatial substructure in the galaxy populations (see, e.g., ~reviews by Girardi \\& Biviano 2002 and Buote 2002) A potentially powerful new tracer of the assembly history of clusters is intracluster light, the diffuse starlight which permeates many galaxy clusters. Once simply another curiosity of Zwicky (1951), individual intracluster stars have been clearly detected in several nearby galaxy clusters \\citep{1996arna,ftv1998,durr2002,ipn3}, and through deep imaging this diffuse light has been detected in many more distant clusters \\citep{uson1991b,vg1994,bern1995,gon2000}. From the results to date, it is clear that intracluster light (ICL) is a common component of galaxy clusters and contains between 10\\% and 50\\% of the total stellar luminosity of the cluster, albeit with large uncertainties due to the intrinsic low surface brightness of the component (less than 1\\% of the night sky background). The study of intracluster light is now entering a new phase, focusing on what can be learned about the evolution of galaxy clusters and their member galaxies using the ICL. Since galaxy clusters form hierarchically, and since the bulk of ICL production is believed to occur due to tidal-stripping from galaxy interactions and from the mean tidal field of the cluster \\citep{rich1983,miller1983,mer1984,gnedin2003}, the properties of the ICL should be intimately tied to the dynamical evolution of clusters (\\eg ~Merritt 1984; Moore \\etal 1996; Dubinski 1998; Mihos 2003; Napolitano \\etal 2003; Willman \\etal 2004). As clusters dynamically evolve, the fractional amount of ICL should increase if the cluster is isolated, and its spatial and dynamical structure should become well-mixed in the cluster potential well. As additional groups of galaxies enter the cluster, the fractional amount of ICL will briefly decrease, as the newer galaxies are initially unstripped, and then the fraction will increase as the additional galaxies suffer the effects of the cluster environment. The exact details of this overall evolution depend on the mass and accretion history of the cluster, tying the properties of the ICL directly to the cosmic history of cluster assembly. The exact mechanisms of ICL production have implications for a number of other galaxy cluster studies. Ultra-compact dwarf galaxies (Drinkwater \\etal 2003), the formation of S0 galaxies (Quilis, Moore, \\& Bower 2000, and references therein) and tidal debris in nearby clusters (Trentham \\& Mobasher 1998; Gregg \\& West 1998; Calcaneo-Roldan \\etal 2000) may all be closely related to the ICL phenomenon. Since searches for individual intracluster stars in nearby galaxy clusters can be influenced by metallicity effects (Durrell \\etal 2002; Feldmeier \\etal 2004), knowing the dominant progenitor population is critical to avoid underestimating the true fraction of intracluster light. However, observational constraints on the ICL are still extremely poor, due largely to the scarcity of quantitative measurements of the ICL in clusters, especially over a range of cluster properties. To address these questions, we have recently begun deep imaging of a sample of galaxy clusters to quantify the structure of ICL as a function of galaxy cluster properties (Feldmeier \\etal 2002; hereafter Paper~I). We have observed two cD-dominated (Bautz-Morgan Type I) galaxy clusters thus far (Abell~1413 and MKW~7); in each we quantify the extended cD envelope and find an excess of luminosity over a pure \\rquart law. However, this extended light is very smooth, and we find relatively little substructure in the ICL of either Abell~1413 or MKW~7. Again, however, both of these clusters are cD-dominated, and there are a number of reasons to expand our sample beyond cD-dominated galaxy clusters. First, less than twenty percent of the total number of Abell clusters have a Bautz-Morgan type of I \\citep{leir1977}. Observations of ICL in Type~I clusters may therefore not be representative of ICL structure in galaxy clusters as a whole. Second, the relation between intracluster light and cD envelopes is still unclear. Does the presence of a cD galaxy in a cluster always imply a large amount of intracluster starlight, or are ICL properties less correlated to the precise details of the cluster core? If intracluster stars are predominantly removed early in the cluster's dynamical history (Merritt 1984), we would expect the properties of ICL to be intimately tied to the process of cD formation. If instead the bulk of the intracluster stars are removed from their parent galaxies by late tidal-stripping and related ``galaxy harassment'' scenarios operating on low luminosity galaxies (Richstone \\& Malumuth 1983; Moore \\etal 1996), then the ICL properties may be less sensitive to the presence or absence of a cD galaxy in the cluster core. Finally, given the expected evolution of galaxy clusters under hierarchical structure formation models (\\eg\\ Merritt 1985; Dubinski 1998), it is expected that most rich galaxy clusters will eventually form cD galaxies, though the timescales involved could be quite long (up to thousands of Gyr; Adams \\& Laughlin 1999). Therefore Bautz-Morgan type II or III clusters may be less dynamically evolved proxies of the more evolved type I clusters, and ICL studies of these clusters may give us insight into to the evolution of galaxy clusters at higher redshift and younger dynamical ages. ", "conclusions": "We have surveyed four Abell type II-III (non-cD) galaxy clusters, and find a significant amount of intracluster light in all of them. The amount of intracluster starlight is intermediate, approximately 10--20\\% of the galaxy light. This may reflect the dynamical adolescence of these galaxy clusters: as the clusters continue to evolve the mean intracluster fraction should increase. However, it is likely that cluster mass (or richness) also plays a role: both our own measurements of the rich cluster Abell~1413 (Paper~I), and other studies of richer clusters such as Coma (Bernstein \\etal 1995) find significantly higher intracluster star fractions than the clusters studied in this paper. Clearly, more observations of clusters over a range of cluster richnesses will be important to separate the two related effects. We have searched for signs of asymmetries and tidal debris in the intracluster light, and we find multiple examples of these features, though in some clusters we found signs of a regular symmetric structure in the intracluster light and few tidal features. The number of tidal features in Abell~1914 is clearly unusual, in both number and surface brightness. It is worth reiterating here that Abell~1914 has a radio halo\\citep{gio1999,kemp2001} and is believed to be undergoing a cluster merger (Jones \\etal 2001). Therefore, the wealth of high surface brightness tidal features in the ICL in Abell~1914 is likely to be due to the ongoing cluster merger. We may be observing the cluster in {\\it flagrante delicto}: had we observed the cluster a few crossing times later, the evidence of such a strong merger would be much less apparent. Comparing our observational results to modern numerical simulations of galaxy clusters, we find good overall qualitative agreement in both the amount and distribution of the intracluster light, although the simulations show that significant ICL substructure may exist well below out current detection levels. From the results found here, we suggest that intracluster light may act as a dynamical ``clock'' of galaxy clusters, one that is complementary to other studies of galaxy clusters such as X-ray substructure, weak lensing, and galaxy radial velocities. However, the comparison of ICL properties and models of structure formation and evolution is still in its infancy. A detailed comparison of the two will require much deeper ICL imaging and measurements of the kinematics of the ICL (though planetary nebulae velocities, for example), as well as through numerical simulations that have larger dynamical range, and better descriptions of galaxy formation processes. Such studies are now underway." }, "0403/astro-ph0403287_arXiv.txt": { "abstract": "We present spectral analysis of the Crab Nebula obtained with the {\\it Chandra} X-ray observatory. The X-ray spectrum is characterized by a power-law whose index varies across the nebula. The variation can be discussed in terms of the particle injection from the pulsar in two different directions: the equatorial plane containing the torus and the symmetry axis along the jet. In the equatorial plane, spectra within the torus are the hardest, with photon index $\\alpha \\approx 1.9$, and are almost independent of the surface brightness. At the periphery of the torus, the spectrum gradually softens in the outer, lower surface brightness regions, up to $\\alpha \\approx 3.0$. This indicates that synchrotron losses become significant to X-ray emitting particles at the outer boundary of the torus. We discuss the nature of the torus, incorporating information from observations at other wavelengths. Spectral variations are also seen within the southern jet. The core of the jet is the hardest with $\\alpha \\approx$ 2.0, and the outer sheath surrounding the core becomes softer with $\\alpha$ up to 2.5 at the outermost part. Based on the similarity between the spectra of the jet core and the torus, we suggest that the electron spectra of the particles injected from the pulsar are also similar in these two different directions. The brightness ratio between the near and far sides of the torus can be explained by Doppler boosting and relativistic aberration; however, the observed ratio cannot be derived from the standard weakly magnetized pulsar wind model. We also found a site where an optical filament comprised of supernova ejecta is absorbing the soft X-ray emission ($<$ 2 keV). ", "introduction": "OBSERVATION AND ANALYSIS} Owing to the X-ray brilliance of the Crab Nebula, observations with the {\\it Chandra} ACIS instrument are technically challenging. The high count rate generally causes two problems in the data acquisition: telemetry saturation and event pile-up (caused by more than one photon depositing charge in a given pixel during a CCD exposure, thereby compromising our ability to measure the energy of each individual photon). Although telemetry saturation results in loss of observing efficiency, it does not lead to biased estimates of physical parameters (e.g., count rate) because events are discarded in a unit of frames, not event by event. On the other hand, pile-up makes interpretation of spectral results problematic, since it converts two photons into a single ``event'' which has apparent energy equal to the sum of the photon energies. This results in underestimated count rate and hardens the detected spectrum, as discussed in Weisskopf et al.\\ (2000). We observed the Crab Nebula eight times (Table~\\ref{tbl:log}), using an instrumental mode of ACIS-S different from that of Weisskopf et al.\\ (2000) in order to reduce pile-up. We adopted a frame time of 0.2 sec instead of the standard frame time of 3.2 seconds, but we did not utilize the grating to reduce count rate. This reduced the counts pixel$^{-1}$ frame$^{-1}$ by a factor of 2.3 with respect to the data of Weisskopf et al.\\ (2000), while improving statistics by increasing the total number of counts by one order of magnitude for a similar effective exposure time. Since the Crab Nebula is a diffuse source, the event pile-up was suppressed by more than a factor of 2.3. The observational parameters (e.g., telemetry format of graded mode, no dithering) were the same through the eight observations. A combined image from the last seven observations is shown in Figure~\\ref{fig:index_map}a (the first observation did not include the whole remnant due to an instrument setup error). This image is an order of magnitude deeper than the image of Weisskopf et al.\\ (2000). In order to study X-ray spectral variations within the Crab Nebula, we divided the whole image into $2.^{\\!\\!\\prime\\prime}5 \\times 2.^{\\!\\!\\prime\\prime}5$ square regions, from which spectra were extracted. The eight data sets were combined in the analysis in order to provide each region with enough statistics. We excluded regions dominated by scattered photons (less than 2000 counts per analysis region) and/or by trailing (or ``out of time'') events. Trailing events, which are caused by photons detected during the CCD readout, can be seen not only as the line emanating from the pulsar along the detector readout direction (which is almost east-west for all of our observations), but also as diffuse emission which dominates the faint regions to the east and west of the nebula due to the enormous brilliance of the Crab PWN. The number of the trailing events in a given pixel can be estimated by multiplying the total number of events integrated along the readout direction by the ratio of the charge transfer row period to the frame exposure time (this ratio is $1.89 \\times 10^{-4}$). We exclude regions where more than 20\\% of the total events recorded in that region are due to the trailing events. We fitted the 2074 spectra with a power-law model in the 0.5--8 keV energy band. We included galactic absorption with a fixed $N_{\\rm H}$ of $3.2 \\times 10^{21} $cm$^{-2}$, which is derived from a fit of the faintest region where pile-up is negligible. The recently discovered quantum efficiency decay of ACIS\\footnote{See http://asc.harvard.edu/cal/Acis/Cal\\_prods/qeDeg/index.html} is taken into account\\footnote{See http://www.astro.psu.edu/users/chartas/xcontdir/xcont.html}. Figure~\\ref{fig:correlation_plot}a shows a correlation plot of apparent photon index versus observed surface brightness. The statistical errors of the apparent photon indices are quite small. They can be roughly expressed empirically as $0.01 \\times I^{-0.5}$ where $I$ is the observed surface brightness in unit of counts s$^{-1}$ arcsecond$^{-2}$ (90\\% confidence level). In spite of the reduction in counts per frame, many of the spectra are still distorted by event pile-up. The spectral parameters derived from a fit of the whole nebula spectrum with a power-law plus galactic absorption model are summarized in Table~\\ref{tbl:whole_fit} (uncorrected values), as are the canonical values for comparison. The pile-up results in a lower apparent spectral index and lower apparent intensity than the canonical values. The observed intensity normalization is about 84\\% of that estimated from the canonical values with the instrumental response. Pile-up has the effect of reducing photon index in regions of increased surface brightness, which is the same general trend seen in the data of Figure~\\ref{fig:correlation_plot}a. The pile-up effects must be corrected in order to determine the nature of any real spectral variations across the Crab nebula. In order to estimate how pile-up distorts the incident spectrum, we utilized the spectral simulator tool LYNX (Chartas et al.\\ 2000). LYNX traces the propagation of individual photons through the mirror and ACIS via the raytrace tool MARX (Wise et al.\\ 1997) and the PSU ACIS Monte Carlo CCD simulator (Townsley et al.\\ 2001), respectively. Since LYNX takes into account the possible overlap of the resulting charge clouds within each CCD exposure frame, spectra affected by pile-up can be simulated for both point sources and diffuse sources. We simulated the observed Crab spectrum for several different values of true surface brightness and for true photon indices of 1.5, 2.0, and 2.5. The results are superposed on the data in Figure~\\ref{fig:correlation_plot}a. The expected reduction of apparent photon index in higher surface brightness regions is evident. However, it is also clear that the variations in observed photon indices greatly exceed those expected from pile-up, and there are certainly real spectral variations across the Crab Nebula, as described by Weisskopf et al.\\ (2000). We used the LYNX simulation results to correct the data points in Figure~\\ref{fig:correlation_plot}a for the photon index and surface brightness errors induced by the pile-up effects. For each data point, we interpolated the simulated data points to the observed surface brightness and photon index and calculated the corresponding true surface brightness and photon index. Figure~\\ref{fig:correlation_plot}b presents a correlation plot of the corrected data points. This simulation was intended to give corrections on the data points in Figure~\\ref{fig:correlation_plot}a and was not adjusted to correct the whole nebula spectrum in order to match to the canonical spectral parameters. However, we used the corrected photon indices and surface brightnesses of each data point in Figure~\\ref{fig:correlation_plot}b to calculate a corrected spectrum for the whole nebula and we present the results in Table~\\ref{tbl:whole_fit} (corrected values). The photon index and the normalization factor of the corrected spectrum agree well with the canonical values within the errors, which strongly supports the success of our pile-up corrections. Hereafter, we use the terms ``photon index'' and ``surface brightness'' to mean these corrected values. We should note a limitation of our simulation. Our corrections are based on modeling of a uniform diffuse source. We apply them to each $2.^{\\!\\!\\prime\\prime}5 \\times 2.^{\\!\\!\\prime\\prime}5$ square region under the reasonable assumption that the surface brightness within each region is uniform. Therefore, application of this technique to small complicated structures like the inner ring or the relativistically expanding wisps requires more complex modeling which is beyond the scope of this work. The systematic errors of the photon index mainly come from this assumption. Judging from the results of the simulation shown in Figure~\\ref{fig:correlation_plot}a and typical surface brightness fluctuations of $\\sim 10\\%$ within an analysis region, the systematic errors in photon index are $\\sim$ 0.05 and dominate the statistical errors. Figure~\\ref{fig:index_map}b shows a map of the photon index. The structure in the torus is more symmetrical about the pulsar in the photon index map than in the broad-band image (Fig.~\\ref{fig:index_map}a). While Doppler boosting and relativistic aberration brighten the northwestern portion of the torus (Pelling et al.\\ 1987) in the Figure~\\ref{fig:index_map}a image, by contrast there is relatively little variation of the photon index within the torus. The hardest structures in the nebula, with photons indices as low as 1.8, are the inner ring and portions of the torus, including the circular structures seen at each extremity of the torus discovered in the first {\\it Chandra} observation (Weisskopf et al.\\ 2000). However, these regions contain small complex structures and the hardest photon index should be re-examined using data set free from pile-up. The entire torus is quite hard ($\\alpha \\approx 1.9$) compared with the outer portions of the nebula. The southern jet is also relatively hard ($\\alpha \\approx 2.0$), whereas the northern counter jet is significantly softer ($\\alpha \\approx 2.25$). The bright region around the counter jet to the northwest of the torus, hereafter called the ``umbrella'' because of its shape, has even softer emission ($\\alpha \\approx 2.5$). Photon indices as large as 3.0 are found in the outer peripheral portions of the nebula. ", "conclusions": "SUMMARY} We have shown spatial variations of the X-ray spectrum of the Crab Nebula in terms of photon index, at an angular scale of arcseconds. The variations can be viewed in two different directions of the particle injection from the pulsar. Across the equatorial plane, the spectrum is almost constant to the outer boundary of the torus, with photon index $\\alpha$ $\\approx$ 1.9 regardless of the surface brightness. It softens significantly, up to $\\alpha$ $\\approx$ 3.0, in the outer, fainter peripheral region. This seems qualitatively consistent with the previous suggestions that the outer boundary of the torus is interpreted as a synchrotron burn-off boundary where the synchrotron losses become significant to X-ray emitting particles. However, the fact that structures similar to the torus are seen at other wavelengths indicates that the torus is not a simple result of synchrotron burn-off. Within the southern jet, photon index variations are also seen: the spectral softening takes place from the central core to the outer sheath. The photon index at the central core is almost the same as that of the torus. This indicates that the electron spectra are similar in the two different directions of the particle injection from the pulsar. We also found that the volume emissivities of the jet and the torus are similar. Assuming that the brightness difference between the near and far sides of the torus is caused by Doppler boosting and relativistic aberration, the ratio can be explained by the observed speed of the downstream flow at the torus. However it cannot be obtained from the so-called weakly magnetized pulsar wind with $\\sigma \\approx 0.003$, as suggested by Kennel \\& Coroniti (1984). Finally, we found that an optical filament comprised of supernova ejecta surrounding the pulsar wind nebula is absorbing the soft X-ray emission from a small portion of the X-ray nebula." }, "0403/astro-ph0403552_arXiv.txt": { "abstract": " ", "introduction": "Chandra observatory, thanks to its sub-arcsec angular resolution, opened a new era in studying X-ray binary populations in nearby galaxies. For the first time an opportunity was presented to observe compact sources in a nearly confusion free regime. The long suspected fact has been proved, that X-ray binaries are an important, if not dominant, contributor to the X-ray emission of the normal galaxies,\\cite{fabbiano2003} as illustrated by the example of our Galaxy.\\cite{grimm02} Depending on the mass of the optical companion, X-ray binaries are subdivided in to two classes -- low and high mass X-ray binaries, having significantly different evolutionary time scale, $\\sim 10^{6-7}$ and $\\sim 10^{9-10}$ years respectively \\cite{xrbrev}. The nearly prompt emission of HMXBs makes them a potentially good tracer of the recent star formation activity in the host galaxy \\cite{rs78}. The LMXBs, on the other hand, have no relation to the present star formation, but, rather, are related to the stellar content of the host galaxy \\cite{lmxb}. Chandra observations of the nearby galaxies presented a possibility to confirm this simple picture and to calibrate the HMXB--SFR\\cite{grimm03,lx-sfr,ranalli} and LMXB$-{\\rm M_*}$\\cite{colbert04,lmxb,kim2004} relations. An unusual class of compact sources -- ultraluminous X-ray sources, has been discovered in nearby galaxies more than a decade ago \\cite{colbert99,fab89}. Although bright, $L_X>10^{39}$ erg/s, point-like sources are found both in young star forming galaxies and in old stellar population of elliptical and S0 galaxies, the most luminous and exotic objects are associated with actively star forming galaxies. Their nature and relation to more ordinary X-ray binaries is still a matter of a significant debate. Based on a simple Eddington luminosity argument, they appear to be powered by accretion onto an intermediate mass object -- a black hole with the mass in the hundreds-thousands solar masses range\\cite{fabbiano2003,miller2003}. However, a number of alternative models have been considered as well -- from collimated radiation\\cite{koerding2002} to $\\sim$stellar mass black holes, representing the high mass tail of the standard stellar evolution sequence and accreting in the near- or slightly super-Eddington regime\\cite{king2001}. \\begin{figure} \\centerline{\\hbox{ \\resizebox{0.45\\hsize}{!}{\\includegraphics{lf_all.ps}} \\resizebox{0.45\\hsize}{!}{\\includegraphics{lf_all_corr.ps}} }} \\caption{{\\em Left:} The XLFs of compact X-ray sources in nearby star forming galaxies. {\\em Right:} The XLFs of the same galaxies normalized to the star formation rates (from Grimm et al.\\cite{grimm03}). The XLF of the Milky Way includes only HMXBs. } \\label{fig:lf_hmxb} \\end{figure} \\begin{figure} \\centerline{\\hbox{ \\resizebox{0.45\\hsize}{!}{\\includegraphics{clf_all_annulus_unsc.ps}} \\resizebox{0.45\\hsize}{!}{\\includegraphics{clf_all_annulus_scaled_mass.ps}} }} \\caption{{\\em Left:} The XLFs of compact X-ray sources in nearby elliptical, S0 galaxies and bulges of spiral galaxies. {\\em Right:} The same XLFs normalized to the stellar mass of the host galaxy (from Gilfanov\\cite{lmxb}). The XLF of the Milky Way includes only LMXBs. } \\label{fig:lf_lmxb} \\end{figure} ", "conclusions": "" }, "0403/astro-ph0403078_arXiv.txt": { "abstract": "A population of black holes (BHs) at high redshifts ($z\\gsim 6$) that contributes significantly to the ionization of the intergalactic medium (IGM) would be accompanied by the copious production of hard ($\\gsim 10$ keV) X-ray photons. The resulting hard X--ray background would redshift and be observed as a present--day soft X--ray background (SXB). Under the hypothesis that BHs are the main producers of reionizing photons in the high--redshift universe, we calculate their contribution to the present--day SXB. Our results, when compared to the unresolved component of the SXB in the range 0.5-2 keV, suggest that accreting BHs (be it luminous quasars or their lower--mass ``miniquasar'' counterparts) did not dominate reionization. Distant miniquasars that produce enough X--rays to only partially ionize the IGM to a level of at most $x_e\\sim 50\\%$ are still allowed, but could be severely constrained by improved determinations of the unresolved component of the SXB. ", "introduction": "\\label{sec:intro} The recent discovery of the Gunn--Peterson (GP) troughs in the spectra of $z>6$ quasars in the Sloan Digital Sky Survey (SDSS; White et al. 2003; Wyithe \\& Loeb 2004) suggests that the end of the reionization process occurs at a redshift near $z\\sim 6$. At this epoch, the ionizing sources drive a strong evolution of the neutral fraction of the intergalactic medium (IGM) from values near unity down to $x_{\\rm HI}\\sim 10^{-3}$ (e.g., Fan et al. 2002). On the other hand, the high electron scattering optical depth, $\\tau_e=0.17 \\pm 0.04$, measured recently by the {\\it Wilkinson Microwave Anisotropy Probe (WMAP)} experiment (Spergel et al. 2003) suggests that ionizing sources were abundant at a much higher redshift, $z\\sim 15$. These data imply that the reionization process is extended and complex, and is probably driven by more than one population of ionizing sources (see, e.g., Haiman 2003 for a post-{\\it WMAP} review). The exact nature of these ionizing sources remains unknown. Natural candidates to account for the onset of reionization at $z\\sim 15$ are massive, metal--free stars that form in the shallow potential wells of the first collapsed dark matter halos (Wyithe \\& Loeb 2003a; Cen 2003a; Haiman \\& Holder 2003). The completion of reionization at $z\\sim 6$ could then be accounted for by a normal population of less massive stars that form from the metal--enriched gas in larger dark matter halos present at $z\\sim 6$. The most natural alternative cause for reionization is the ionizing radiation produced by gas accretion onto an early population of black holes (``miniquasars''; see Haiman \\& Loeb 1998, Wyithe \\& Loeb 2003c, Bromm \\& Loeb 2003). The ionizing emissivity of the known population of quasars diminishes rapidly beyond $z\\gsim 3$, and bright quasars are unlikely to contribute significantly to the ionizing background at $z\\gsim 5$ (Shapiro, Giroux \\& Babul 1994; Haiman, Abel \\& Madau 2001; Wyithe \\& Loeb 2003a). However, if low--luminosity, yet undetected miniquasars are present in large numbers, they could dominate the total ionizing background at $z\\sim 6$ (Haiman \\& Loeb 1998). Recent work, motivated by the {\\it WMAP} results, has emphasized the potential significant contribution to the ionizing background at the earliest epochs ($z\\sim 15$) from accretion onto the seeds of would--be supermassive black holes (Madau et al. 2003; Ricotti \\& Ostriker 2003). The soft X--rays emitted by these sources can partially ionize the IGM early on (Oh 2001; Venkatesan \\& Shull 2001). A population of miniquasars at $z\\gsim 6$ would be accompanied by the presence of an early X-ray background. Since the IGM is optically thick to photons with energies $E$ below $E_{\\rm max} = 1.8 [(1+z)/15)]^{0.5} x_{\\rm HI}^{1/3}\\hs {\\rm keV}$, the soft X-rays with $E\\lsim E_{\\rm max}$ would be consumed by neutral hydrogen atoms and contribute to reionization. However, the background of harder X-rays would redshift without absorption and would be observed as a present--day soft X--ray background (SXB). In this paper, {\\it we examine the hypothesis that accreting BHs are the main producers of reionizing photons in the high--redshift universe, and calculate their contribution to the present--day SXB in this case}. Our results, when compared to the unresolved component of the SXB, suggest that accreting BHs cannot contribute significantly either to the completion of reionization at $z > 6$, or to the partial ionization of the IGM at $z\\gsim 15$ to ionized fractions of $x_e\\gsim 0.5$. The outline of the paper is as follows: In \\S~\\ref{sec:method}, we describe the method to calculate the SXB from quasars that contribute to reionizing the universe. In \\S~\\ref{sec:sxrb}, we critically discuss current X-ray observations, focusing on the unresolved fraction of the SXB that could be attributed to distant quasars. In \\S~\\ref{sec:fullion}, we calculate the expected contribution to the SXB from hypothetical quasars and their lower--mass miniquasar counterparts that fully ionized the universe. In \\S~\\ref{sec:preion}, we repeat our analysis for a putative miniquasar population that partially ionizes the IGM at high redshifts. In \\S~\\ref{sec:discuss}, we discuss how various simplifications made in our analysis influence our final results. Finally, in \\S\\ref{sec:conclusions} we summarize our results and the implications of this work. Throughout this paper, we adopt the background cosmological parameters as measured by the {\\it WMAP} experiment, $\\Omega_m=0.27$, $\\Omega_{\\Lambda}=0.73$, $\\Omega_b=0.044$, and $h=0.71$ (Spergel et al. 2003) and set the mass fraction of helium to $Y_{\\rm He}=0.24$ ( e.g. Burles, Nollett, \\& Turner, 2001). In the rest of the paper `accreting BHs' will refer to both quasars and their lower--mass ``miniquasar'' counterparts. \\newpage ", "conclusions": "\\label{sec:conclusions} Our main results, obtained in \\S~\\ref{sec:quasars} and \\S~\\ref{sec:miniquasars}, are summarized in Table~\\ref{table2}. The third column in this table denotes the percentage of the total SXB contributed by accreting BHs (which we used to refer to both quasars and their lower--mass ``miniquasar'' counterparts) for models in which high redshift accreting BHs contribute significantly to full or partial reionization. These percentages can be compared directly with the percentages given by M03. The second column gives the total intensity of the quasar populations, to be compared to the allowed range of the mean or maximum unresolved SXB flux of $0.35 \\times 10^{-12}$ \\unitE to $1.23 \\times 10^{-12}$ \\unitE, respectively. As Table~\\ref{table2} shows, models in which $z>6$ accreting BHs contribute to reionization overproduce the SXB. Pre-ionization by miniquasars requires fewer ionizing photons, both because only a fraction of the H atoms need to be ionized, and because hard X--rays can produce multiple secondary ionizations. We find that models in which X--rays are assumed to partially ionize the IGM up to $x_e \\sim 0.5$ at $6\\la z\\la 20$ are still allowed, but could be severely constrained by improved determinations of the unresolved component of the SXB. We emphasize that our constraints derive from the total number of ionizing photons that the population as a whole needs to produce. Therefore, our conclusions depend mostly on the assumed spectral shape, and are independent of the details of the population, such as the luminosity function and its evolution with redshift. Future improvements in resolving the SXB, improving the limits on the unresolved component by a factor of a few, would place stringent constraints on the contribution of $z\\sim 15$ accreting BHs to the scattering optical depth measured by \\wmap." }, "0403/astro-ph0403234_arXiv.txt": { "abstract": "We explore the self-enrichment hypothesis for globular cluster formation with respect to the star formation aspect. Following this scenario, the massive stars of a first stellar generation chemically enrich the globular progenitor cloud up to Galactic halo metallicities and sweep it into an expanding spherical shell of gas. This paper investigates the ability of this swept proto-globular cloud to become gravitationally unstable and, therefore, to seed the formation of second generation stars which may later on form a globular cluster. We use a simple model based on a linear perturbation theory for transverse motions in a shell of gas to demonstrate that the pressures by which the progenitor clouds are bound and the supernova numbers required to achieve Galactic halo metallicities support the successful development of the shell transverse collapse. Interestingly, the two parameters controling the metallicity achieved through self-enrichment, namely the number of supernovae and the external pressure, also rule the surface density of the shell and thus its ability to undergo a transverse collapse. Such a supernova-induced origin for the globular cluster stars opens therefore the way to the understanding of the halo metallicity distributions. This model is also able to explain the lower limit of the halo globular cluster metallicity. ", "introduction": "Globular clusters (GC) are dense, massive and round-shaped groups of stars present in the vast majority of galaxies. In our Galaxy, the halo GCs were among the very first bound structures to form and their study provides therefore valuable information about the early Galactic evolution. Their formation is an exciting but yet unsolved problem. For instance, it is still an open question whether Galactic halo GCs formed out of gas already chemically enriched ({\\it pre}-enrichment models, e.g. Harris \\& Pudritz 1994) or whether they produced their own heavy elements through an earlier generation of stars within the GC progenitor itself ({\\it self}-enrichment models). In the second class of models, the issue of their formation is directly related to the origin of their metal content (Cayrel 1986; Brown, Burkert \\& Truran 1995; Parmentier et al.~1999). Such a feature makes the self-enrichment scenario especially appealing if we hypothesize that the GC progenitor clouds are made of primordial (i.e. metal-free) gas, a reasonable assumption for the Old Halo GCs, i.e. the population of old and coeval (Rosenberg et al.~1999) halo GCs. Regarding the origin of the proto-globular cluster clouds (PGCC), Fall \\& Rees (1985) suggested that they formed out of the collapsing protoGalaxy, as cold and dense clouds in pressure equilibrium with a hot and diffuse background. In the frame of this theory, the GC progenitor clouds are thermally supported (i.e. no additional support against gravitation due to magnetic fields and turbulence) and made of primordial gas. In order to explain the metallicities of halo GCs, Parmentier et al.~(1999, hereafter Paper I) expanded the Fall \\& Rees (1985) model for GC formation by the self-enrichment picture. According to this one, a first stellar generation forms in the central regions of each PGCC. When the massive stars explode as Type II supernovae (SNeII), they chemically enrich the surrounding gas and sweep the cloud, turning it in an expanding shell of gas in which the formation of a second, chemically enriched, stellar generation may be triggered. These second generation stars form the proto-globular cluster. \\\\ The sweeping and compression of the interstellar medium by massive star explosions is not the sole mechanism invoked to account for star formation during the earliest stages of the Galactic evolution. Some other models assume a different origin for the trigger. For instance, following Vietri \\& Pesce (1995), the high pressure confining the GC gaseous progenitor leads to the propagation of a strong shock inwards the cloud, stimulating thereby the formation of new stars. On another side, Murray \\& Lin (1992) and Dinge (1997) suggested that the propagation of shock waves is promoted by cloud-cloud collisions. In a similar way, star forming clouds may have coalesced into larger units until they reach a density and/or accumulated mass high enough to enable the formation of bound globular clusters (Larson 1988, Smith 1999). Obviously, several processes are able to collect ambient gas into dense layers/clouds where star formation can thereafter take place. Several of them may have been at work at the same time. Our interest being in understanding the origin of the metal content of GCs, in what follows, we investigate the hypothesis of star formation in gas layers swept by the explosions of PopIII massive stars located at the center of the PGCCs. \\\\ The debate of how PopIII stars looked like has been raging for many years. Numerous studies have addressed the issue of the collapse and fragmentation of primordial gas clouds in order to estimate the masses of PopIII stars. The primordial gas being deficient in heavy elements (i.e., the most efficient coolants in present-day star forming clouds), all of them have emphasized the importance of cooling by H$_2$ molecules. Inspite of this, the achieved conclusions do not necessarily converge. Recent numerical simulations (Abel, Bryan \\& Norman 2002) suggest that metal-free stars form in isolation and are massive (30 $\\lesssim$ M $\\lesssim$ 100~M$_{\\odot}$) objects. From their own simulations, Bromm, Coppi \\& Larsson (1999) quoted a characteristic mass even larger than 100~M$_{\\odot}$. These results are in marked contrast with the early study performed by Palla, Salpeter \\& Stahler (1983) following which primordial gas clouds should be capable of fragmenting into low-mass stars (i.e., down to $\\sim$ 0.1~M$_{\\odot}$). Nakamura \\& Umemura (1999) reached a somewhat intermediate conclusion. According to them, the mass range of the first stars is similar to its present value (i.e. no star with M $\\gtrsim$ 100~M$_{\\odot}$) but nevertheless excludes low-mass long-lived stars, that is, the lowest mass star allowed to form in a metal-free medium is a $\\simeq$ 3~M$_{\\odot}$ star. In a more recent model, Nakamura \\& Umemura (2001) predict a bimodal initial mass function for PopIII stars. In fact, the initial mass function of metal-free stars would show two distinct peaks at $\\simeq$ 1~M$_{\\odot}$ and $\\simeq$ 100~M$_{\\odot}$, that is, it would include intermediate mass (1 $\\lesssim$ M $\\lesssim$ 10~M$_{\\odot}$) stars, massive (10 $\\lesssim$ M $\\lesssim$ 100~M$_{\\odot}$) stars as well as very massive (M $\\gtrsim$ 100~M$_{\\odot}$) ones. As quoted by Christlieb et al.~(2003), the recent discovery of HE 0107-5240, the most metal-poor ([Fe/H]=$-$5.3) star ever discovered in the Galactic halo, could be a challenge to these models precluding the formation of stars with mass low enough for their life duration to exceed a Hubble time. While there is currently a wide consensus that present star formation operates predominantly in a clustered mode (Lada, Strom \\& Myers 1993), the question whether clusters of metal-free stars managed to form in the first (proto-) galaxies is still open. In this paper, we assume that at least some of the primordial star formation sites produced stellar {\\it clusters} whose most massive stars (10 $\\lesssim$ M $\\lesssim$ 50\\,M$_{\\odot}$) end their life as canonical SNeII. Therefore, our study does not include the possibility of a prompt enrichment of the primordial interstellar medium by a population of very massive (i.e., $\\gtrsim$ 100~M$_{\\odot}$) stars as suggested by, e.g., Wasserburg \\& Qian (2000). The chemical enrichment provided by isolated (very) massive objects will be qualitatively discussed in Sect.~3 where we will see that they may be appropriate to account for the formation of very metal-poor stars, i.e., stars more metal-poor than the most metal-deficient halo GCs ([Fe/H] $\\lesssim$ $-$2.5). \\\\ Supernovae having long been thought to disrupt the cloud of gas out of which they have formed, Parmentier et al.~(1999) studied the ability of pressure-truncated clouds to retain SNII ejecta. This ability will rule the metal content of the proto-cluster. In fact, in this class of models, the final metallicity is determined by the number of supernovae ($N$, which, assuming a given initial mass function and given SNII yields, determines the amount of metals dispersed within the PGCC) and the background pressure ($P_h$, which determines the mass of the cloud, that is, the mass of primordial gas to be chemically enriched). Comparing the gravitational energy of the PGCC with the kinetic energy of the supershell resulting from the SNII explosions, Parmentier et al.(1999) showed that the gaseous GC progenitors can sustain up to $\\sim$ 200 supernovae (the disruption criterion, Eq.~14, Paper I). This is a number high enough for the PGCCs to achieve halo metallicities. Furthermore, for a given number of exploding massive stars, a self-enrichment episode in pressure-bound clouds lead to a metallicity gradient throughout the resulting system of GCs and to a correlation between the mass and the achieved metallicity of the gaseous progenitors in the sense that the least massive clouds are the most metal-rich. Such a trend emerges because if the bound pressure is higher, the mass of the pressure-truncated cloud will be lower, and its ability to retain supernova ejecta will be greater. These trends, i.e. a metallicity gradient and a mass-metallicity relation, are indeed statistically present in the Old Halo, that is, the halo from which the presumably accreted component has been removed (Parmentier et al.~2000, Parmentier \\& Gilmore 2001). \\\\ After having analysed to which extent the Galactic halo GC data fit the correlations induced by the self-enrichment process, the next step is to wonder whether there are some stars forming out of the chemically enriched supershell, i.e. whether there is a second stellar generation tracing these correlations. The idea of star formation induced by supernova explosions dates back at least to Opik (1953). Numerous cases of distant star forming loops, connected with shells resulting from supernova explosions and located in the Galactic disc as well as in dwarf galaxies, are detailed in the literature (e.g. Comeron \\& Torra 1994, Walter et al. 1998, Efremov \\& Elmegreen 1998, Rubio et al. 1998). In this paper, we address the specific case of supershells made of swept PGCCs and within which the formation of the stars of future GCs may be triggered. In a first step, we limit our study to the propagation of the shell throughout the hot protogalactic background in which the PGCCs are initially embedded, this part of the shell propagation being much longer than the propagation throughout the cloud (see Fig.~\\ref{fig:RsHPB}). \\\\ The outline of the paper is as follows. In Sect.~2, we solve the perturbed equations of continuity and motion for transverse flows within the shell (i.e. the swept cloud) in order to identify the conditions supporting a successful shell transverse collapse. We also discuss in turn the impact of the different parameters acting upon the shell collapse and, thus, upon the temporal growth of the shell fragments in which further star formation may be stimulated. In Sect.~3, we show how the conditions required to stimulate a star formation episode within the shell provides a natural explanation to the observed metallicity range of halo GCs and how the shape of their metallicity spectrum constitutes the next step to work on. Sect.~4 describes some effects which our forthcoming computations should take into account in order to refine the present model. Finally, our conclusions are presented in Sect.~5. \\section[]{Stimulated Star Formation in Proto-Globular Cluster Clouds} \\label{sec:stimSF} The shell will in general contain perturbed (transverse) velocity components and perturbations in the column density whose development leads to the transverse collapse of the swept PGCC and, thereby, to the formation of a second stellar generation. We now derive the conditions to get such a collapsing shell. The elementary method described below is adopted. \\subsection{Modelling the transverse collapse of the shell} The computations are based on the linear perturbed equations of continuity and motion for transverse flows in the shell (e.g. Elmegreen 1994). \\\\ The perturbed equation of continuity (mass conservation) is \\begin{equation} \\frac{\\partial \\sigma _1}{\\partial t} = -2 \\frac{V_s}{R_s} \\sigma _1 - \\sigma _0 \\nabla _T . v\\,, \\label{eq:per_cont} \\end{equation} where subscript T means that the gradient component under consideration is the transverse one, and the perturbed equation of motion (momentum conservation) is \\begin{equation} \\sigma _0 \\frac{\\partial v}{\\partial t} = - \\sigma _0 \\frac{V_s}{R_s} v - {c_s}^2~ \\nabla \\sigma _1 + \\sigma _0 g_1\\;. \\label{eq:per_motion} \\end{equation} In these equations, $R_s$ and $V_s$ are respectively the radius and the velocity of the shell, $\\sigma _0$ is the unperturbed surface density, $\\sigma _1$ is the perturbed surface density, $v$ is the perturbed (transverse) velocity, $c_s$ is the velocity dispersion of the material inside the shell, $g_1$ is the perturbed gravity, this one being related to the surface density through (Elmegreen 1994): \\begin{equation} g _1 = -2 \\pi i G \\sigma _1\\;. \\label{g1_sigma1} \\end{equation} As mentioned above, the evolution of the shell is studied while propagating through the hot background, i.e. when the whole cloud has been swept inside the shell. The hot background being a diffuse medium, the shell mass $M_s(t)$ does not increase any longer at this stage of its propagation and is given by the mass $M$ of the progenitor cloud. The unperturbed surface density of the shell is thus given by \\begin{equation} \\sigma _0 = \\frac{1}{4 \\pi} \\frac{M_s(t)}{R_s ^2} = \\frac{1}{4 \\pi} \\frac{M}{R_s ^2}\\;. \\label{sigma_0} \\end{equation} Equation \\ref{eq:per_cont} shows that the development with time of any perturbation of the shell surface density ($\\partial \\sigma _1 / \\partial t > 0$) is inhibited by the stretching of the perturbed region due to the shell expansion (i.e. $V_s > 0$, first term on the right hand-side, hereafter rhs) while the convergence of the perturbed flows supports the growth of the perturbation (second term on the rhs). Equation \\ref{eq:per_motion} shows that an initial transverse flow of material along the shell develops ($\\partial v/ \\partial t > 0$) only if the self-gravity (third term on the rhs) overcomes the stabilizing effects of the stretching (first term on the rhs) and of the internal pressure (second term on the rhs), here represented by $c_s^2$, the shell sound speed squared. \\\\ In order to solve Eqs.\\ref{eq:per_cont} and \\ref{eq:per_motion} properly, we now turn to the determination of the expansion law of the shell, i.e. $R_s(t)$ and $V_s(t)$, while it propagates throughout the hot protogalactic background. \\subsection{Supershell propagation throughout the hot background} \\label{sub:propa_HPB} \\begin{figure} \\begin{center} \\epsfig{figure=RsHPB.eps, width=\\linewidth} \\caption{Evolution with time of the shell radius inferred from Eqs.~(\\ref{eq:bubble_nrj_HPB} - \\ref{eq:mass_sh_HPB}) assuming that 200\\,SNeII explode at a constant rate during 30 million years and for three different hot protogalactic background pressures, from top to bottom 5$\\times 10^{-11}$ (plain curve), 10$^{-10}$ (dashed-dotted curve) and 5$\\times 10^{-10}$\\,dyne.cm$^{-2}$ (dotted curve). In each case, an arrow indicates the time $t_{em}$ at which the shell crosses the interface between the cold and hot phases} \\label{fig:RsHPB} \\end{center} \\end{figure} The equations describing the propagation of a supernova-driven shell are as follow (Castor et al.~1975). \\begin{enumerate} \\item The supernova explosions add energy to the bubble at a constant rate $\\dot{E_o}$ and the dominant energy loss of the bubble comes from the work against the dense shell, hence the variation with time of the energy $E_b$ of the bubble obeys \\begin{equation} \\dot{E_b}=\\dot{E_o}-4\\pi {R_s}^2 P_b \\dot{R_s}\\,. \\label{eq:bubble_nrj_HPB} \\end{equation} We assume that the kinetic energy of every SN is 10$^{51}$~ergs and that the SN phase lasts about thirty million years, i.e. $\\dot{E_o}=N 10^{51} {\\rm ergs}/30 {\\rm Myr}$. \\item The internal energy $E_b$ and the pressure $P_b$ of the bubble are related through \\begin{equation} \\frac{4\\pi}{3} {R_s}^3 P_b = \\frac{2}{3} E_b\\,, \\label{eq:bubble_P_HPB} \\end{equation} \\item The shell motion obeys Newton's second law \\begin{equation} \\frac{d}{dt} [M_s(t) \\dot{R_s}(t)] = 4\\pi {R_s}^2 (P_b-P_{ext}) -\\frac{{GM_s}^2(t)}{2{R_s}^2(t)}\\,, \\label{eq:Newton2_HPB} \\end{equation} where $P_{ext}$ is the pressure of the medium just outside the shell. \\item Considering the case of swept PGCCs propagating through the hot protogalactic background, the mass of the shell is constant in time and is given by the mass of the cloud (see Sect.~2.1): \\begin{equation} M_s(t) = M\\;. \\\\ \\label{eq:mass_sh_HPB} \\end{equation} \\end{enumerate} The pressure external to the shell is exerted by the surrounding hot background (i.e. P$_{ext}$ = P$_h$ in Eq.~\\ref{eq:Newton2_HPB}) and is therefore assumed to be constant in time. This pressure will strongly decelerate the shell as illustrated below. Numerical resolution of Eqs.~(\\ref{eq:bubble_nrj_HPB} - \\ref{eq:mass_sh_HPB}) provides $R_s(t)$ and thereby $V_s(t)$ and $\\sigma _0(t)$. The initial conditions are those at the time $t_{em}$, i.e. when the shell crosses the interface between the cloud and the surrounding background: the mass and radius of the shell are those of the pressure-bound cloud ($R_s(t_{em})=R$, $M_s(t_{em})=M$), the velocity of the shell is determined by its former propagation at constant speed $V$ through the cloud (see Paper I, Eq.~13), that is $V_s=V$ and $A_s=0$. \\\\ \\begin{figure} \\begin{center} \\epsfig{figure=RsHPBt_3.eps, width=\\linewidth} \\caption{Comparison between the evolution with time of the shell radius computed from Eqs.~(\\ref{eq:bubble_nrj_HPB} - \\ref{eq:mass_sh_HPB}) and the average radius given by Eq.~9 (i.e. $ \\propto t^{1/3}$) for the quoted numbers $N$ of SNeII and background pressures $P_h$} \\label{fig:RsHPBt1/3} \\end{center} \\end{figure} Figure \\ref{fig:RsHPB} shows the evolution with time of the shell radius for 200\\,SNeII, i.e. the maximum number of supernovae that a PGCC can sustain (disruption criterion, Paper I), and for 3 different values of the hot protogalactic background pressure, i.e. $P_h$ = 5$\\times 10^{-11}$, 10$^{-10}$ and 5$\\times 10^{-10}$\\,dyne.cm$^{-2}$. The SN rate being the same in the three cases, the shells propagate at the same velocity in the cold phase (Eq.~13 in Paper I). Their expansions begin to differ once they have crossed the interface between their respective PGCC and the hot background in which the latter is embedded, this time being indicated by an arrow in Fig.~\\ref{fig:RsHPB}. Obviously, the propagation of the shell through the cloud is much shorter than its propagation through the background. A significant part of the shell expansion through the background takes place at early time. Indeed, after a transient phase during which the velocity of the shell does not differ markedly from its velocity inside the cloud (Fig.~\\ref{fig:RsHPB}), the overall expansion slows down and the radius of the shell scales roughly as $t^{1/3}$. This average expansion can be obtained from Eqs.~(\\ref{eq:bubble_nrj_HPB} - \\ref{eq:mass_sh_HPB}) assuming that $M_s(t)=0$, reflecting thereby that the temporal evolution of the shell radius does not depend strongly on the mass (Brown et al.~1995): \\begin{equation} = \\left (\\frac{3}{10\\pi} \\frac{\\dot{E_o}}{P_h} \\right )^{1/3} t^{1/3}. \\label{eq:RsHPBt1/3} \\end{equation} Figure \\ref{fig:RsHPBt1/3} shows the good agreement between Eq.~\\ref{eq:RsHPBt1/3} and the result of the numerical integration over time of Eqs.~(\\ref{eq:bubble_nrj_HPB} - \\ref{eq:mass_sh_HPB}). Equation \\ref{eq:RsHPBt1/3} shows that, during the long-term evolution, the expansion rates of two shells having the same $(N/P_h)$ ratio are similar. Figure \\ref{fig:RsHPBt1/3} illustrates this effect for two sets of values, namely $N$=100 and $P_h$=5$\\times 10^{-11}$\\,dyne.cm$^{-2}$ (plain curve) and $N$=200 and $P_h$=$10^{-10}$\\,dyne.cm$^{-2}$ (dashed-dotted curve). \\subsection{Growth with time of the shell fragments} \\label{sec:frag_num} In order to assess whether the shell transverse collapse proceeds successfully or not, we now numerically integrate over time Eqs.\\ref{eq:per_cont} and \\ref{eq:per_motion} in order to derive the temporal evolutions of the perturbed and unperturbed surface densities, $\\tilde{\\sigma} _1(t)$ and $\\sigma _0(t)$, respectively. Assuming that the perturbed quantities follow a complex exponential of the angular position $\\phi$ along the shell, we get: \\begin{equation} \\sigma _1 (t,\\phi) = \\tilde{\\sigma} _1 (t) ~e^{-i \\eta \\phi} \\label{sigma1_exp} \\end{equation} and \\begin{equation} v (t,\\phi) = \\tilde{v} (t) ~e^{-i \\eta \\phi} ~e^{i \\Delta \\phi}\\,. \\label{v_exp} \\end{equation} In these equations, $\\Delta \\phi$ represents the phase difference between the perturbed surface density $\\sigma _1$ and the perturbed velocity $v$. $\\eta$ is the angular wavenumber and is related to the spatial wavenumber $k$ by: \\begin{equation} \\eta = k R_s = \\frac{2 \\pi}{\\lambda}~R_s \\label{eta_lambda} \\end{equation} where $\\lambda$ is the wavelength of the perturbation, namely the average distance between forming fragments, the sites of future star formation. Therefore, $\\eta$ is the number of forming clumps along a shell circumference and, as such, $\\eta$ must be an integer. Moreover, any realistic perturbation must fit inside a fraction of the shell circumference, say, $\\lambda \\leq R_s$ or $\\eta \\geq 6$. \\\\ At a time $t$ and an angular position $\\phi$ along the shell, the shell surface density $\\sigma _s$ obeys \\begin{equation} \\sigma _s(t,\\phi )=\\sigma _0(t)+\\sigma _1(t,\\phi ) =\\sigma _0(t)+\\tilde{\\sigma _1}(t)~cos(\\eta \\phi)\\;. \\label{eq:sigt_t_anlt} \\end{equation} The fragmentation of the shell, i.e.~its fully developed transverse collapse, occurs when \\begin{equation} \\tilde{\\sigma} _1(t) = \\sigma _0(t). \\end{equation} Writing $- i \\eta /R_s$ for the gradient and using Eqs.~\\ref{g1_sigma1} and \\ref{sigma_0}, Eqs.~\\ref{eq:per_cont} and \\ref{eq:per_motion} successively become: \\begin{itemize} \\item[$\\triangleright$] Perturbed equation of continuity: \\begin{eqnarray} \\frac{\\partial \\tilde{\\sigma _1}}{\\partial t} & = & -2\\frac{V_s}{R_s} ~\\tilde {\\sigma _1} + \\sigma _0 ~ \\frac{i~\\eta}{R_s} ~ \\tilde{v}~e^{i\\Delta \\phi} \\nonumber \\\\ \\frac{\\partial \\tilde{\\sigma _1}}{\\partial t} & = & -2\\frac{V_s}{R_s} ~\\tilde{\\sigma _1} + \\frac{i~\\eta~M}{4 \\pi} \\frac{\\tilde{v}}{{R_s}^3} ~e^{i\\Delta \\phi} \\label{eq:per_cont_num} \\end{eqnarray} \\item[$\\triangleright$] Perturbed equation of motion: \\begin{eqnarray} \\sigma _0 ~e^{i\\Delta \\phi} ~\\frac{\\partial \\tilde{v}}{\\partial t} & = & -\\sigma _0 ~ \\frac{V_s}{R_s} ~ \\tilde{v}~e^{i\\Delta \\phi} + {c_s}^2 ~ \\frac{i~\\eta}{R_s} ~ \\tilde{\\sigma _1} \\nonumber \\\\ & & \\,- 2 \\pi i G ~ \\sigma _0 ~ \\tilde{\\sigma _1} ~~~ \\nonumber \\\\ ~\\frac{\\partial \\tilde{v}}{\\partial t} & = & - \\frac{V_s}{R_s} ~ \\tilde{v} \\hspace{1.3cm} \\,+ \\frac{4~\\pi ~i ~\\eta ~c_s^2}{M} R_s ~\\tilde{\\sigma _1} ~ e^{- i\\Delta \\phi} \\nonumber \\\\ & & - 2 \\pi i G \\tilde{\\sigma _1} e^{- i\\Delta \\phi} ~~~ \\label{eq:per_motion_num} \\vspace*{3pt} \\end{eqnarray} \\end{itemize} Seven parameters intervene in the ability of the shell to undergo a transverse collapse: \\begin{itemize} \\item[- ] the supernova number $N$ and the pressure external to the shell $P_h$, which determine the shell expansion law ($R_s$ depends on $N$ and $P_h$) and mass ($M$ depends on $P_h$), \\item[- ] the initial values of the perturbed quantities, $\\tilde \\sigma _1(t_{em})$ and $\\tilde v (t_{em})$, and the associated phase difference $\\Delta \\phi$, \\item[- ] the number $\\eta$ of forming clumps embedded along a shell circumference, \\item[- ] the velocity dispersion $c_s$ of the shell material which is related to the internal pressure $P_s$ of the shell, supporting it against transverse collapse, through $P_s={c_s}^2 \\rho _s$ ($\\rho _s$ is the volumic mass density of the shell material). \\end{itemize} In what follows, the influence of each of the parameters involved in the fragmentation process is investigated, the final aim being to check whether some reasonable sets of conditions can lead to the shell fragmentation. \\subsubsection{The phase difference between $v$ and $\\sigma_1$: $\\Delta \\phi$} \\label{sub:delta_phi} \\begin{figure} \\hspace*{5mm} \\epsfig{figure=frag_shell.eps, width=11cm} \\caption{Schematic description of the sinusoidal (i.e. complex exponential) behaviour with $\\phi$ of the perturbed quantities $v$ (lower curve) and $\\sigma _1$ (upper curve) along a fraction of the shell circumference at a given time assuming that the number of clumps $\\eta$ is 15 and the phase difference $\\Delta \\phi$ between $v$ and $\\sigma _1$ is $-\\frac{\\pi}{2}$. Such a phase difference corresponds to the convergence of the transverse (perturbed) flows (indicated by the thick arrows) towards the initial clumps (filled circles). The open circles represent the shell regions which are progressively depleted by the transverse flows} \\label{fig:fragshell} \\end{figure} The amplitude of the perturbed surface density at a time $t$, $\\tilde \\sigma_1(t)$, results from its initial value $\\tilde \\sigma_1(t_{em})$ and from the transverse flows which redistribute the shell mass accumulated while the shell was propagating throughout the progenitor cloud. The fragmentation is therefore favoured if these transverse flows converge towards the clumps initially present within the shell. Such a situation corresponds to a phase difference $\\Delta \\phi$ = $-\\frac{\\pi}{2}$ between the two perturbed quantities $\\sigma _1(t,\\phi)$ and $v(t,\\phi)$, that is, the transverse velocity exhibits a phase delay of a quarter of a wavelength with respect to the perturbed surface density (see Fig.~\\ref{fig:fragshell}). Replacing $\\Delta \\phi$ by this value, Eqs.~\\ref{eq:per_cont_num} and \\ref{eq:per_motion_num} respectively become: \\begin{equation} \\frac{\\partial \\tilde{\\sigma _1}}{\\partial t}=-2\\frac{V_s}{R_s} ~\\tilde {\\sigma _1} + \\frac{\\eta~M}{4 \\pi} \\frac{\\tilde{v}}{{R_s}^3} \\label{eq:per_cont_num_phi}\\,, \\end{equation} \\begin{equation} \\frac{\\partial \\tilde{v}}{\\partial t} =-~\\frac{V_s}{R_s} ~ \\tilde{v} - \\frac{4~\\pi ~\\eta ~c_s^2}{M} R_s ~\\tilde{\\sigma _1} + 2\\pi G ~ \\tilde{\\sigma _1}\\;. \\\\ \\label{eq:per_motion_num_phi} \\end{equation} In what follows, the roles of the other parameters is studied assuming that $\\Delta \\phi = - \\pi $/2, i.e. through the numerical integration of Eqs.~\\ref{eq:per_cont_num_phi} - \\ref{eq:per_motion_num_phi}. \\subsubsection{The global parameters: $N$ and $P_h$} \\label{sub:discNPh} In order to assess whether the shell fragments or not during its propagation through the hot protogalactic background, we compare in Fig.~\\ref{fig:disc_sig1init} the evolutions with time of the perturbed and unperturbed surface densities, namely $\\tilde{\\sigma} _1(t)$ and $\\sigma _0(t)$, for different values of the external pressure $P_h$ (5$\\times 10^{-10}$ and 10$^{-10}$\\,dyne.cm$^{-2}$) and numbers $N$ of SNeII (100 and 200). The corresponding metallicity is indicated in each panel. The other parameters are kept the same in every case: the sound speed $c_s$ of the shell material is 1\\,km.s$^{-1}$, the perturbed velocity $v(t_{em})$ is 0.01\\,$V_s(t_{em})$, the number of forming clumps $\\eta$ is 10 and the initial perturbed surface density is assumed to be of order one per cent of the unperturbed value, namely $\\sigma _1(t_{em}) = 0.01 \\times \\sigma _0(t_{em})$. This choice may seem rather arbitrary but the next paragraph will show that the initial amplitude of $\\sigma _1$ does not affect the results significantly . \\\\ While the unperturbed surface density decreases due to the shell expansion at constant mass, the perturbed surface density grows at a rate which depends on the external pressure and on the number of SNeII. The comparison of the different panels in Fig.~\\ref{fig:disc_sig1init} shows that the development of $\\tilde{\\sigma} _1(t)$ with respect to $\\sigma_0(t)$ is favoured by a {\\sl a high background pressure} and a {\\sl low number of SNeII}. Among the three panels presented ($P_h=10^{-10}$\\,dyne.cm$^{-2}$ and $N$=100, $P_h=5\\times 10^{-10}$\\,dyne.cm$^{-2}$ and $N$=100, $P_h=5\\times 10^{-10}$\\,dyne.cm$^{-2}$ and $N$=200), the transverse collapse of the shell is successfully achieved\\footnote{At this point, it should be kept in mind that once the shell has achieved its complete transverse collapse, Eqs.~\\ref{eq:per_cont} and Eqs.~\\ref{eq:per_motion} are no longer valid. Therefore, once ${\\sigma} _0 \\leq \\tilde\\sigma _1$, the curve $\\tilde \\sigma _1(t)$ has no longer physical meaning.} when the lower number of supernovae is combined with the larger external pressure ($P_h=5\\times 10^{-10}$\\,dyne.cm$^{-2}$ and $N$=100). This effect comes from the dependence of the shell surface density on the external pressure $P_h$ and the supernova number $N$. Indeed, combining the mass of a pressure-truncated gas cloud (i.e. $M \\propto P_h^{-1/2}$) with Eq.~\\ref{eq:RsHPBt1/3}, the surface density of the shell scales as: \\begin{equation} \\sigma _0 (t) \\propto \\frac{M}{R_s(t)^2} \\propto \\frac{P_h^{1/6}}{N^{2/3}}. \\label{eq:sig0_N_Ph} \\end{equation} Consequently, the larger the pressure and/or the lower the supernova number, the larger the shell surface density and the larger the growth rate of the perturbation (see Eq.~\\ref{eq:per_cont} and Eq.~\\ref{eq:per_motion}, a larger surface density favours the terms promoting an efficient transverse collapse). {\\sl Therefore, although the presence of exploding massive stars is required to stimulate the formation of a second stellar generation, too large a number of SNeII inhibits the ability of the shell to collapse and to form new stars}\\footnote{Unlike the protogalactic shells we study, the collapse of shells expanding in the Galactic disc is made easier by a larger number of SNeII. In fact, assuming that the surrounding interstellar medium is roughly homogeneous, i.e. the shell radius is smaller than the density scale-height of the Galactic HI layer, one gets by mass conservation, assuming a pre-shell density $\\rho _0$, $\\sigma _0 = \\frac{R_s \\rho _0}{3}$. In this case, a larger number of SNeII leads to a larger radius at a given time and, therefore, to larger surface density and perturbation growth rate.}. The influence of the surface density of the shell on its collapse is reminiscent of the star formation law on large scales, i.e. averaged over entire galactic discs. In that case, providing that the gas surface density is larger than a density threshold, the star formation rate follows a power-law of the gas surface density (the so-called Schmidt law) while it falls sharply below (Kennicutt 1989). The threshold surface density varies from one galaxy to another but remains nevertheless in the range 10$^{20}$-- 10$^{21}$\\,cm$^{-2}$. It may not be a coincidence that the shell surface densities obtained in the frame of this model are of the same order of magnitude or even larger (see Sect.~2.3.3). The dependence of the perturbation growth rate on $N$ and $P_h$ raises a point of interest, worthy of a mention here. It just so happens that the parameters determining the final metallicity of the proto-cluster, namely the number of supernovae ($N$, which determines the amount of metals dispersed within the PGCC) and the background pressure ($P_h$, which determines the mass of the cloud, that is, the mass of primordial gas to be chemically enriched) are also some of those influencing the ability of the shell to form new stars. {\\sl If some combinations of external pressures and SN numbers support the completion of the shell transverse collapse much more than some others do, then the formation of second stellar generations with the corresponding metallicities will be favoured}. Therefore, in the frame of the self-enrichment scenario, there is a direct link between the achieved metallicity and the probability of forming halo stars, i.e. {\\sl the study of the fragmentation process may shed light on the metallicity distribution function of Galactic halo field stars and GCs}. Figure \\ref{fig:disc_sig1init} and Eq.~19 show that the shell fragmentation is favoured by a low number of SNe and a high pressure of the hot protogalactic background. Now, let us imagine that, on the contrary, the collapse efficiency increases with both decreasing number of SNeII and pressure. As $P_h$ and $N$ get smaller, the number of successful transverse collapses increases and the newly formed stars are more metal-poor. Thus, the metallicity distribution function of the Galactic halo would exhibit an increasing number of proto-clusters/halo stars with decreasing metallicity. If, on the other hand, the shell ability to achieve fragmentation increased with both increasing external pressure and SN number, then the larger $P_h$ and $N$, the more numerous the transverse collapses and the more metal-rich the newly formed stars. As a result, there would be an increasing number of proto-clusters/halo stars with increasing metallicity. Neither an increasing nor a decreasing metallicity distribution function is observed for the Galactic halo. In contrast, the metallicity distributions, for both halo field stars (Laird et al.~1988) and halo GCs (Zinn 1985), are peaked-shape. While the finding that the shell transverse collapse is favoured by large external pressures (promoting ``large'', i.e. mildly metal-poor, metallicities) and low SN numbers (supporting low metallicities) is not sufficient in itself to draw some definitive conclusions regarding the shape of the halo metallicity distribution, it appears that it does not contradict it either. \\subsubsection{The initial perturbed surface density: $\\tilde{\\sigma} _1(t_{em})$} \\label{sub:disc_sig1init} \\begin{figure} \\epsfig{figure=discNPh_4.eps, width=\\linewidth} \\epsfig{figure=discNPh_5bis.eps, width=\\linewidth} \\epsfig{figure=discNPh_6.eps, width=\\linewidth} \\caption{Temporal evolution of $\\sigma _0$ (plain curves) and $\\tilde \\sigma _1$ (dashed-dotted curves), expressed in units of 1\\,M$_{\\odot}$.pc$^{-2}$, considering two different initial perturbed surface densities (see the keys) and the external pressures and SNII numbers indicated. In each panel, $c_s$=1\\,km.s$^{-1}$, $\\eta=10$ and $v(t_{em})=0.01\\,V_s(t_{em})$} \\vspace{11pt} \\vspace{11pt} \\label{fig:disc_sig1init} \\end{figure} Protogalactic shells such as those studied here exhibit number surface densities of order several 10$^{20}$\\,cm$^{-2}$. Indeed, using Eqs.~\\ref{sigma_0} and \\ref{eq:RsHPBt1/3}, one gets: \\begin{eqnarray} \\sigma _0(t, N, P_h) & \\simeq & 5 \\times 10^{-3} P_{h(10)}^{1/6} N_{200}^{-2/3} t_6^{-2/3}\\, {\\rm g.cm^{-2}} \\nonumber \\\\ & \\simeq & 2.5 \\times 10^{21} P_{h(10)}^{1/6} N_{200}^{-2/3} t_6^{-2/3}\\,{\\rm cm^{-2}} \\end{eqnarray} where the subscript (10) means that the pressure is expressed in units of $10^{-10}\\,\\rm dyne.cm^{-2}$, the subscript 200 means that $N$ is expressed in units of 200\\,SNe and the subscript 6 means that the time is expressed in units of $10^6$\\,years. Considering the interstellar medium of the Galactic disc with roughly the same surface density (though a bit lower, i.e. $\\simeq$ 10$^{20}$\\,cm$^{-2}$), the relative non-uniformities in the number surface density are typically of order 0.01 - 0.2 (Wunsch \\& Palous 2001). However, the initial perturbation in the shell surface density is certainly not as large as some of the clumps of this {\\sl quiescent} disc interstellar medium since {\\sl turbulent mixing} is expected to take place within the shell (this is an important requirement to achieve the efficient mixing of the supernova heavy elements with the cloud gas and, therefore, to maintain the chemical homogeneity of the shell, that is, of the proto-globular cluster; see Brown et al.~1991), thus lowering the surface density inhomogeneities. \\\\ Figure 4 also shows the temporal evolution of $\\sigma _0$ and $\\tilde \\sigma _1$ assuming that the initial perturbed surface density is lower than in Sect.~2.3.2, namely 0.25\\%\\,$\\sigma _0(t_{em})$ instead of 1\\%\\,$\\sigma _0(t_{em})$. Despite these different initial values, each panel displays almost identical $\\tilde \\sigma _1(t)$ curves in both cases. Owing to the initial transverse flows, the initial clumps quickly undergo a replenishment in shell material leading afterwards to very similar temporal evolutions of the perturbed surface density whatever the initial perturbed surface density. Obviously, this one is not a key parameter of the fragmentation process. \\subsubsection{The initial perturbed velocity: $v(t_{em})$} \\label{sub:disc_vtrans} All the results presented here above assume the spherical symmetry of the supershell. It is clear however that such a system is not expected to remain perfectly spherical. The deviations from spherical symmetry can arise, for instance, from the non point-like nature of the energy input, namely all the stars of the first generation cluster are not located exactly at the centre of the cloud/shell. A rough estimate of the initial amplitude of the transverse motions within the shell can therefore be derived from the size of this cluster. A star located at a distance $r$ from the shell centre will induce a transverse velocity $v$ such that: \\begin{equation} v ~\\simeq ~V_s ~ \\frac{r}{R_s}\\;. \\vspace{11pt} \\label{eq:vt_1} \\end{equation} In order to estimate the size $r$ of the cluster of massive stars hosted by a PGCC, we refer to R\\,136, the dense core of the 30\\,Doradus Nebula located in the Large Magellanic Cloud. The 30\\,Doradus nebula shows an impressive example of a two-stage stellar formation. The energetic activity of a very compact bright cluster, R136, which includes several tens of O stars, triggers the formation of a new stellar generation revealed by numerous infrared sources in or near some bright filaments west and northeast of R136 (Rubio et al.~1998). Based on Hubble Space Telescope photometry, Campbell et al.~(1992) detected about 160 stars more massive than 10\\,M$_{\\odot}$ in R\\,136 which they define as a region of 2.2\\,pc $\\times$ 1.9\\,pc. This number of massive stars being remarkably similar to the numbers of SNeII used in our self-enrichment model, we adopt R\\,136 as the most similar example in the Local Group of what may have been the first generation cluster. A radius of 1\\,pc appears therefore as a reasonable estimate of the size of this cluster, the source of the energy input. \\\\ Considering an average background pressure of 10$^{-10}$\\,dyne.cm$^{-2}$ and the corresponding cloud radius ($\\simeq$ 30\\,pc, see Eq.~3, Paper I), the transverse velocity when the shell reaches the cloud boundary is: \\begin{equation} v(t_{em}) ~\\simeq ~\\frac{1\\,pc}{30\\,pc} ~V_s(t_{em}) ~\\simeq 0.03 V_s(t_{em})\\;. \\label{eq:vt_2} \\end{equation} Figure 5 displays the evolution with time of $\\sigma _0$ and $\\tilde \\sigma _1$ assuming three different values for the initial transverse velocity, namely 0.01\\,$V_s(t_{em})$, 0.02\\,$V_s(t_{em})$ and 0.03\\,$V_s(t_{em})$, while keeping all the other parameters to their previous values. In sharp contrast with $\\tilde \\sigma _1(t_{em})$, $\\tilde v(t_{em})$ appears to be an important parameter of the shell transverse collapse. For instance, considering the bottom panel in Fig.~5 ($P_h=5 \\times 10^{-10}$\\,dyne.cm$^{-2}$ and $N$=200), the fragmentation takes place less than 15 million years after the first SN explosion if $v(t_{em})=0.03 \\,V_s(t_{em})$, whereas it is prevented if $v(t_{em})=0.01 \\,V_s(t_{em})$. \\begin{figure} \\epsfig{figure=discNPh_12.eps, width=\\linewidth} \\epsfig{figure=discNPh_11bis.eps, width=\\linewidth} \\epsfig{figure=discNPh_10.eps, width=\\linewidth} \\caption{Temporal evolution of $\\sigma _0$ (plain curves) and $\\tilde \\sigma _1$ (dashed-dotted curves), expressed in units of 1\\,M$_{\\odot}$.pc$^{-2}$, considering three initial perturbed velocities (see the keys) and the external pressures and SNII numbers indicated. In each panel, $c_s$=1\\,km.s$^{-1}$, $\\eta$=10 and $\\tilde \\sigma _1(t_{em})=0.01\\,\\sigma _0 (t_{em})$} \\vspace{11pt} \\vspace{11pt} \\label{fig:disc_vtrans} \\end{figure} \\subsubsection{The number of forming clumps: $\\eta$} \\label{sub:disc_eta} \\begin{figure} \\epsfig{figure=discNPh_7.eps, width=\\linewidth} \\epsfig{figure=discNPh_8bis.eps, width=\\linewidth} \\epsfig{figure=discNPh_9.eps, width=\\linewidth} \\caption{Temporal evolution of $\\sigma _0$ (plain curves) and $\\tilde \\sigma _1$ (dashed and dotted curves), expressed in units of 1\\,M$_{\\odot}$.pc$^{-2}$, considering the external pressures, SNII numbers and numbers of forming clumps indicated (see the keys). In each panel, $c_s$=1\\,km.s$^{-1}$, $\\tilde \\sigma _1(t_{em})=0.01\\sigma _0(t_{em})$, $v(t_{em})=0.01\\,V_s(t_{em})$} \\vspace{11pt} \\label{fig:disc_eta} \\end{figure} In supernova-driven shells of gas, star formation does not take place all along the whole periphery but, instead, takes place in regularly spaced clumps (i.e., where the transverse collapse brings some gas from depleted adjoining regions in the shell, see Fig.~\\ref{fig:fragshell}). The number $\\eta$ of such clumps along a shell circumference is thus related to the wavelength $\\lambda$ of the perturbation (Eq.~\\ref{eta_lambda}). This aspect of shell collapse modelling following which stars are formed in discrete stellar subsystems along a shell periphery is supported by several observations of star forming shells, the so-called Sextant being one of the most illustrative. In the Large Magellanic Cloud, a set of five OB associations are located along a HI supershell, sustaining there about 1/6 of a complete circle and being thus called the Sextant (Efremov \\& Elmegreen 1998). Efremov, Ehlerova \\& Palous (1999) showed that the OB associations are regularly spaced, the deprojected average distance between two subsequent stellar groups being $\\sim$ 37~pc. Furthermore, using numerical simulations of shells expanding in a mass model of the Large Magellanic Cloud, they concluded that the formation of these 5 OB associations is most probably the result of a triggered star formation episode in the supershell created by SNeII located near the Sextant centre. Their simulations also explain how projection effects make the visible star forming regions only a fraction of a total circle. \\\\ The influence of the number of clumps embedded within the shell on the transverse collapse is illustrated in Fig.~\\ref{fig:disc_eta}. The parameter $\\eta$ is not as negligeable as the initial perturbed surface density. In some cases, it acts upon the fragmentation process almost as strongly as the initial transverse velocity. It is thus worth estimating the angular wavenumbers which are the most favourable to the growth of an initial perturbation. In order to do so, we now derive an analytical approximation of the instantaneous growth rate of the perturbation. This is done straightforwardly by assuming that the perturbed quantities vary exponentially with time $t$, in analogy with the exponential growth rate found in other instability problems (e.g. the Jeans mass) even though there is no real exponential growth in this problem owing to the shell expansion and the corresponding time dependence of $\\sigma _0$. It is thus important to keep in mind that the perturbation growth rate derived below (Eqs.~\\ref{eq:omega_eta} and \\ref{eq:fg_omega}) is {\\it not} used to compute the temporal evolution of $\\sigma _1$ (the cases displayed in Fig.~\\ref{fig:disc_eta} are obtained from solving Eqs~\\ref{eq:per_cont_num_phi}-\\ref{eq:per_motion_num_phi}) but merely to derive $\\eta$ estimates favouring the shell collapse. The perturbed quantities are written: \\begin{equation} \\sigma _1 (t,\\phi) = \\tilde{\\sigma}_1(t_{em}) ~e^{\\omega t} ~e^{-i \\eta \\phi} \\label{sigma1_exp} \\end{equation} and \\begin{equation} v (t,\\phi) = \\tilde{v}(t_{em}) ~e^{\\omega t} ~e^{-i \\eta \\phi} ~e^{i \\Delta \\phi}, \\label{v_exp} \\end{equation} where $\\omega$ is the angular frequency of the perturbation. Following Eqs.~\\ref{sigma1_exp} and \\ref{v_exp}, we write $\\omega$ for the time derivatives and $- i \\eta/R_s$ for the transverse gradients. Therefore, Eqs.~\\ref{eq:per_cont} and \\ref{eq:per_motion} become respectively \\begin{equation} \\omega \\sigma _1 = -2 \\frac{V_s}{R_s} \\sigma _1 + \\sigma _0 \\frac{i \\eta}{R_s} v \\label{eq:per_cont2} \\end{equation} and \\begin{equation} \\sigma _0 \\omega v = - \\sigma _0 \\frac{V_s}{R_s} v + {c_s}^2~ \\frac{i \\eta}{R_s}\\sigma _1 - 2 \\pi i G \\sigma _0 \\sigma _1 \\,, \\label{eq:per_motion2} \\end{equation} using Eq.~\\ref{g1_sigma1} in the latter. \\\\ The elimination of the perturbed quantities $\\sigma _1$ and $v$ between Eqs.~\\ref{eq:per_cont2} and \\ref{eq:per_motion2} provides the dispersion equation, namely the relation between the angular frequency $\\omega$ (i.e. the instantaneous growth rate) and the angular wavenumber $\\eta$ of the perturbation: \\begin{equation} \\sigma _0 \\left( \\omega + 2 \\frac{V_s}{R_s} \\right) \\left( \\omega + \\frac{V_s}{R_s} \\right) - \\sigma _0 \\frac{i \\eta}{R_s} \\left( 2 \\pi i G \\sigma _0 - c_s^2 \\frac{i \\eta}{R_s} \\right) = 0\\,, \\end{equation} whose solution is given by \\begin{equation} \\omega (\\eta) = -\\frac{3}{2} \\frac{V_s}{R_s} + \\sqrt{\\frac{3}{2} \\frac{V_s ^2}{R_s ^2} + 2 \\pi G \\sigma _0 \\frac{\\eta}{R_s} - c_s^2 \\frac{\\eta ^2}{R_s ^2}}. \\label{eq:omega_eta} \\end{equation} Let us consider the {\\sl first growing mode}. This one corresponds to the sequence of values of $\\eta$ which maximises the angular frequency $\\omega$ at each moment of the shell propagation, i.e. $\\eta _{fg} = \\eta (t)$ such that \\begin{equation} \\frac{d \\omega}{d \\eta} = 0\\;. \\label{eq:fg_def} \\end{equation} Equation \\ref{eq:fg_def} indeed corresponds to a maximum since the discriminant of Eq.~\\ref{eq:omega_eta} shows a negative curvature with $\\eta$. The instantaneous angular wavenumber and the instantaneous angular frequency associated to the first growing mode obey respectively \\begin{equation} \\eta _{fg} = \\frac{\\pi G}{{c_s}^2} \\sigma _0 R_s = \\frac{1}{4 {c_s}^2} \\frac{GM}{R_s} \\label{eq:fg_eta} \\end{equation} and \\begin{equation} \\omega _{fg} = - \\frac{3}{2} \\frac{V_s}{R_s} + \\sqrt{\\frac{V_s^2}{R_s^2} + \\frac{\\pi ^2 G^2 \\sigma _0^2 }{c_s^2} }\\;. \\label{eq:fg_omega} \\end{equation} Taking into account the dependence of $M$ and $R_s$ (Eq.~\\ref{eq:RsHPBt1/3}) on $P_h$ and $N$, Eq.~\\ref{eq:fg_eta} shows that $\\eta _{fg}$ depends on the external pressure $P_h$, on the SN number $N$ and on time $t$ as \\begin{equation} \\eta _{fg} \\propto P_h ^{-1/6} N^{-1/3} t^{-1/3}\\;. \\label{eq:eta_fg/NPh} \\end{equation} A first guess of a favourable angular wavenumber can be estimated from the temporal average of Eq.~\\ref{eq:fg_eta} over the time spent by the supershell in the hot protogalactic background: \\begin{equation} <\\eta _{fg}> = \\frac{1}{\\Delta t - t_{em}} \\int_{t_{em}}^{\\Delta t} \\eta _{fg}(t') ~ dt' \\,, \\label{eta_fg_ave} \\end{equation} where $\\Delta t$ is the duration of the SN phase. Among the cases displayed in Fig.~\\ref{fig:disc_eta} ($c_s$=1\\,km.s$^{-1}$, $\\tilde \\sigma _1(t_{em})=0.01\\sigma _0(t_{em})$, $v(t_{em})=0.01\\,V_s(t_{em})$), we see that the shell transverse collapse is achieved if, for instance, $P_h = 10^{-10}$\\,dyne.cm$^{-2}$, $N$=100 and $\\eta$=16 (top panel) or if $P_h = 5 \\times 10^{-10}$\\,dyne.cm$^{-2}$, $N$=100 and $\\eta$=10 (middle panel). It is interesting to note that these values of $\\eta$ reasonably match those given by Eq.~\\ref{eta_fg_ave}, for the above mentioned cases, i.e., $<\\eta _{fg}>$=12 and $<\\eta _{fg}>$=9, respectively. As Eq.~\\ref{eta_fg_ave} has been derived under the assumption of an exponential growth rate, this agreement a posteriori justifies its validity as a convenient estimate of the number of forming clumps in collapsing shells. \\\\ \\subsubsection{The shell sound speed: $c_s$} \\label{sub:disc_cs} \\begin{figure} \\begin{center} \\epsfig{figure=disc_cs.eps, width=\\linewidth} \\label{fig:disc_cs} \\caption{Temporal evolution of $\\sigma _0$ (plain curve) and $\\tilde \\sigma _1$ (dashed curves), expressed in units of 1\\,M$_{\\odot}$.pc$^{-2}$, considering the sound speeds (see the key), external pressures and SNII numbers indicated. Other parameters are $\\eta$=10, $\\tilde \\sigma _1(t_{em})=0.01\\sigma _0(t_{em})$, $v(t_{em})=0.01\\,V_s(t_{em})$} \\end{center} \\end{figure} The development of a gravitational instability within a supershell also depends on the sound speed $c_s$ of the shell material. Indeed, $c_s$ is directly related to the thermal pressure $P_s$ of the shell gas through \\begin{equation} P_s = \\frac{\\rho _s k T_s}{\\mu _s m_H} = \\rho _s c_s ^2\\,. \\end{equation} In this equation, $k$ and $m_H$ are the Boltzmann constant and the hydrogen mass, respectively, while $\\rho _s$, $T_s$ and $\\mu _s$ are the mass density, the temperature and the mean molecular weight of the shell, respectively. The larger the velocity dispersion, the more the layer resists gravitational collapse. In an H\\,I layer (T $\\simeq$ 100\\,K, $\\mu ~\\simeq$ 1.3\\,m$_H$), the sound speed is $\\simeq$ 0.8\\,km.s$^{-1}$, while it can be as low as $\\simeq$ 0.3\\,km.s$^{-1}$ in an H$_2$ layer (T $\\simeq$ 20\\,K, $\\mu ~\\simeq$ 2.1\\,m$_H$) (McCray \\& Kafatos 1987). However, the turbulence and magnetic fields of the shell will increase these values. To some extent, they can be represented by an additional pressure term in the expression of $c_s$, i.e. \\begin{equation} c_s = \\left ( \\frac{k T_s}{\\mu _s m_H} + \\frac{B_s ^2}{4 \\pi \\rho _s} + \\mathcal{T} {\\rm _s} \\right)^{1/2} \\label{eq:cs_mag} \\end{equation} where $B_s$ is the magnetic field in the shell and $\\mathcal{T} {\\rm _s}$ is the contribution of turbulence. \\\\ Figure 7 shows how highly sensible to $c_s$ the fragmentation process is. Considering 100\\,SNeII and a hot background pressure of 5 $\\times$ 10$^{-10}$\\,dyne.cm$^{-2}$, the fragmentation takes place even with a low initial transverse velocity, i.e. $v(t_{em})=0.01 V_s(t_{em})$, if $c_s$=1\\,km.s$^{-1}$. Increasing the latter by 25\\%, the transverse collapse is very weakened and the fragmentation is prevented. During the last ten million years, the evolutions with time of $\\sigma _0$ and $\\tilde \\sigma _1$ are similar, that is, the evolution of $\\tilde \\sigma _1 (t)$ is mostly driven by the dilution of $\\sigma _0 (t)$ due to the shell expansion. \\\\ Equation \\ref{eq:cs_mag} illustrates the difficulty of estimating the sound speed of a gas. It implies the computations of its cooling history, its magnetic fields and turbulence. Moreover, the high sensibility of the fragmentation issue to $c_s$ (see Fig.~7) shows that its value must be estimated with some accuracy. Such a task is well beyond the scope of the present work and, in what follows, we adopt $c_s=1\\,{\\rm km.s}^{-1}$, in agreement with many studies of supershell fragmentation (e.g. Comeron \\& Torra 1994, Ehlerova \\& Palous 2002). ", "conclusions": "\\label{sec:search_para} \\subsection{Stars with halo GC metallicities} \\label{sec:disc_GC} The previous section has shown that the shell/swept PGCC may become gravitationally unstable and finally break into fragments providing that some conditions are fulfilled, e.g. $v(t_{em})/V_s(t_{em}) \\simeq $ 0.03, $c_s \\simeq $ 1\\,km.s$^{-1}$, $\\eta \\simeq <\\eta _{fg}>$. This is {\\sl not} to claim that all supershells will encounter such favourable circumstances, but one may expect that at least {\\sl some} of them will do. \\\\ \\begin{figure} \\begin{minipage}[b]{\\linewidth} \\begin{center} \\epsfig{figure=frag_v1.eps, width=\\linewidth} \\end{center} \\end{minipage} \\vfill \\vspace*{-9mm} \\begin{minipage}[b]{\\linewidth} \\begin{center} \\epsfig{figure=frag_v2.eps, width=\\linewidth} \\end{center} \\end{minipage} \\vfill \\vspace*{-9mm} \\begin{minipage}[b]{\\linewidth} \\begin{center} \\epsfig{figure=frag_v3.eps, width=\\linewidth} \\end{center} \\end{minipage} \\label{fig:frag_prob} \\caption{Summary of the ($N$, $P_h$, $v(t_{em})/V_s(t_{em})$, $\\eta$) values which may (open or filled circles), or may not (no symbol), lead to a swept PGCC fragmentation providing that the transverse flows converge towards the clumps embedded within the shell, i.e. $\\Delta \\phi = -\\pi /2$. The sound speed and the initial perturbed surface density are $c_s = 1$\\,km.s$^{-1}$ and $\\sigma _1 (t_{em}) = 0.01 ~\\sigma _0 (t_{em})$, respectively. The top, intermediate and bottom panels correspond to initial transverse velocity of 1, 2 and 3 per cent of the shell velocity when it emerges out of the cloud, respectively. The isometallicity curves corresponding to [Fe/H]=$-$1.2, $-$1.5, $-$2, $-$2.5 are also plotted, giving thus the metallicity achieved through self-enrichment for each parameter combination. For each ($N$,$P_h$) pair, 54 couples of initial perturbed velocities ($v(t_{em})/V_s(t_{em}$)= 0.01, 0.02, 0.03) and clump numbers ($\\eta$ ranging from 6 to 40 by step of 2) are tested. In each panel (i.e., for each value of the initial perturbed velocity), a couple ($N$, $P_h$) leading to a successful transverse collapse is marked by an open/filled circle as well as by the range of $\\eta$ values leading to fragmentation. The larger the range of $\\eta$ values, the bigger the symbol (see text for details)} \\end{figure} Figure 8 presents the results of shell fragmentation simulations for varying values of $N$, $P_h$, $\\eta$ and $v(t_{em})/V_s(t_{em})$, assuming that $c_s$=1\\,km.s$^{-1}$ and $\\sigma _1 (t_{em}) $= 0.01 $\\sigma _0(t_{em})$. Five hot protogalactic background pressures ($P_h=10^{-11}, 3.2 \\times 10^{-11}, 10^{-10}, 3.2 \\times 10^{-10}, 10^{-9}$\\,dyne.cm$^{-2}$) and 3 SN numbers ($N$=50, 100, 200) are tested. The upper limit for $N$ is the maximum number of supernovae that the GC gaseous progenitor can sustain, namely $N$=200 (i.e., if $N > 200$, the absolute value of the cloud binding energy is lower than the shell kinetic energy: disruption criterion, Paper I). The lower limit is imposed by $c^{PGCC}$, the sound speed of the PGCC material. Indeed, the shell is ``built'' while sweeping the PGCC and such a mass accumulation into a shell requires the velocity of the shell to be larger than the sound speed of the ambiant medium. Therefore, the lower limit to the shell velocity in the PGCC obeys: \\begin{equation} {V_s}^{PGCC} = {c}^{PGCC} = \\sqrt{\\frac{k T}{\\mu m_H}} = 8.3\\,{\\rm km.s^{-1}} \\end{equation} where $T$ and $\\mu$ are the temperature ($\\simeq 10^4$\\,K) and the mean molocular weight ($\\simeq 1.2$) of the PGCC, respectively (Fall \\& Rees 1985). The number of SNeII corresponding to this lower limit of the shell velocity is $N \\simeq 50$ (Eq.~13, Paper I). Regarding the upper value of $P_h$, we refer to Murray \\& Lin (1992) who showed that the hot protogalactic background pressure depends on the galactocentric distance $D$ as \\begin{equation} P_{h} = 1.25\\times 10^{-9} D_{kpc}^{-2} \\,\\, \\rm dyne\\,cm^{-2}. \\label{Ph2} \\end{equation} Thus, $P_h$ is of order 10$^{-9}$\\,dyne.cm$^{-2}$ in the very inner Galactic regions (i.e., $D \\simeq$ 1\\,kpc). \\\\ In order to see what is the metallicity achieved in the shells which succeed in forming new stars, iso-metallicity curves corresponding to [Fe/H]=$-$1.2, $-$1.5, $-$2 and $-$2.5 (i.e., metallicities typical of the Galactic halo GCs) are displayed in each panel of Fig.~8. \\\\ For each couple ($N$,$P_h$), 54 combinations of $v(t_{em})/V_s(t_{em})$ and $\\eta$ (i.e. 3 initial perturbed velocities $\\times$ 18 numbers of clumps) have been run. The initial transverse velocities correspond to 1, 2 and 3 per cent of the velocity of the shell when it enters the hot background and the number of clumps ranges from 6 to 40 by step of two. It appears that the shell is unable to fragment if $\\eta \\ge 40$. When a set of conditions leads to a successful fragmentation, the corresponding point in the ($N$, $P_h$) diagram is marked by a circle as well as by the range of $\\eta$ values leading to successful transverse collapses. Depending on whether the shell fragmentation takes place for 0 to 25 per cent, 25 to 50 per cent, 50 to 75 per cent, or 75 to 100 per cent of the number of $\\eta$ values tested, the corresponding triplet ($N$, $P_h$, $v(t_{em})/V_s(t_{em}$)) is marked by an open circle, a small, a median-size or a large filled circle, respectively. Figure 8 confirms that the probability of successful transverse collapse fades away with decreasing external pressure and increasing number of SNeII. As mentioned in Sect.~\\ref{sub:discNPh}, a low external pressure and a high number of SNeII leads to a larger radius for the shell, decreasing thereby its surface density (Eq.~\\ref{eq:sig0_N_Ph}) and its ability to collapse. \\\\ At high pressure, i.e. $P_h \\simeq 10^{-9}$\\,dyne.cm$^{-2}$, the ability of the shell to get fragmented is limited by too large a number of SNeII. Figure 8 shows however that a SN number as large as 200 does not prevent the fragmentation. Accordingly, the largest metallicity which can be achieved through self-enrichment is [Fe/H]$\\simeq -$1.2. On the other hand, at low pressure, the transverse collapse is supported by a low number of SNeII. Combining a low background pressure with the lower limit on $N$, Fig.~8 shows that the lowest metallicity which can be achieved is [Fe/H]$\\simeq -$2.8 providing that the relative initial transverse velocity is 3 per cent. While this lower limit in metallicity is a bit uncertain, as it is achieved in the case of the highest intial transverse velocity only, the examination of the three panels in Fig.~8 clearly shows that a metallicity of $\\simeq -2.5$ is actually achievable. These extreme values ([Fe/H]$\\simeq -$1.2 and $-$2.5) match nicely the metallicities exhibited by the most metal-rich and the most metal-poor Galactic halo GCs, respectively. \\\\ At this stage, it is worth keeping in mind that the question addressed here above concerns the ability of the shell to form stars and not yet the ability of these stars to evolve into a stellar cluster. We will address the ability of the newly formed stars to form a bound cluster in a forthcoming paper and we emphasize here that, among the cases of successful fragmentation displayed by Fig.~8, some of the newly formed stars may not be able to form a bound cluster. In other words, the collapsed shells may be a source for both halo GCs and halo field stars. Section 2.3.2 has discussed how the self-enrichment model for GC formation can shed light on the metallicity distribution of the Galactic halo. Because the parameters which determine the metallicity, i.e. $N$ and $P_h$, also control the shell surface density, there is a direct link between the metallicity and the probability of star formation\\footnote{This is {\\sl not} to say that the probability of star formation is determined by the metallicity, as it can be assumed in pre-enrichment models in which the metallicity may control the star formation through line cooling processes. It rather means that the metallicity and the probability of star formation derive from the same parameters, namely the number of SNeII and the pressure of the hot protogalactic background}. Furthermore, the initial radius and velocity of the second generation stars being the radius and the velocity of the shell at the time of their formation, their binding depends on the shell expansion law and, therefore, again, on $N$ and $P_h$. As a consequence, the probability of getting a bound cluster from a shell of newly formed stars is also related to their metallicity. If we assume that a significant fraction of the field content comes from ``failed GCs'' (i.e. shells of stars which did not succeed in forming GCs), then these relations between the metallicity on the one hand and the probabilities of forming a second stellar generation and a bound cluster on the other hand {\\sl open the way to the computation of the metallicity distributions of both halo field stars and halo GCs}. Since the formation of a GC requires one more condition than the shell transverse collapse (i.e. the binding of the second stellar generation), only a subset of the ($N$,$P_h$) combinations giving rise to shell fragmentation may lead to the formation of a bound cluster. Such a conclusion does not contradict, and even fits, the fact that the metallicity spectrum of halo field stars is larger than the one of GCs (e.g. Laird et al.~1988). \\\\ \\subsection{The very metal-poor stars ([Fe/H] $\\lesssim -2.5$)} \\label{sub:vmp_st} The most striking difference between the halo GC and halo field metallicity distributions resides in the much more extended metal-poor tail of the field compared to the GC system. In fact, while the most metal-poor GCs show [Fe/H]$\\simeq -$2.5, the Galactic halo hosts field stars whose metallicity is much lower (e.g., CS22876-032, [Fe/H]$\\simeq -$3.7, Norris, Beers \\& Ryan 2000). The recent discovery of HE0107-5240, the most metal-deficient star ever discovered ([Fe/H]$\\simeq -$5.3, Christlieb et al.~2002) has decreased even more the lower limit of the observed metallicity distribution of halo field stars. Our model does not seem to be able to explain these very metal-poor stars (see Fig.~8) and we must therefore investigate alternative scenarios for the formation of these stars more metal-deficient than the most metal-poor GCs. \\subsubsection{Isolated Type II Supernovae} The existence of very metal-poor stars has often been attributed to star formation episodes triggered by {\\it individual} SNeII exploding in primordial gas clouds (e.g., Shigeyama \\& Tsujimoto 1998, Argast et al. 2000, Karlsson \\& Gustafsson 2001), that is, a formation scenario similar to the self-enrichment scenario for GCs except regarding the considered number of massive stars. The metallicity of such stars is determined by the ratio of the mass of metals ejected by the SNII to the mass of hydrogen gathered by the shock wave. The SNII yields computed by Woosley \\& Weaver (1995) for zero metallicity stars show that a SNII with progenitor mass $m$ ejects a mass of metals $m_z$ such that $m_z \\simeq 0.3m - 3.5$ (in unit of one solar mass). On the other hand, the mass $M_{sw}$ of interstellar gas collected by the shock wave is related to the explosion energy $E_0$ through \\begin{equation} M_{sw} = 5.1 \\times 10^4 M_{\\odot} \\left( \\frac{E_0}{10^{51}\\,ergs} \\right)^{0.97}\\,, \\label{single_SN} \\end{equation} assuming a sound speed of 10\\,km.s$^{-1}$ (or T $\\sim$ 10$^4$\\,K) for the interstellar gas (Shigeyama \\& Tsujimoto 1998). Using these relations and assuming $E_0 = 10^{51}$\\,ergs for a canonical SNII, the metallicity achieved in shells of gas driven by isolated SNeII is straightforwardly derived (see the second column of Table 1). The explosion energy being the same whatever the SNII progenitor mass, every SNII sweeps the same amount of primordial circumstellar gas, irrespective of the SNII mass (Eq.~\\ref{single_SN}). Therefore, the final metallicity is determined solely by the mass of heavy elements released in the interstellar medium by the exploding star and, thus, by the SNII progenitor mass. As a consequence, the more massive the SNII, the larger the final metallicity. As already noticed by Shigeyama \\& Tsujimoto (1998), the formation of very metal-poor stars with [Fe/H] $\\simeq -4$ and [Fe/H] $\\simeq -2$ may therefore be ascribed to ``low''-mass (i.e., M $\\simeq$ 12\\,M$_{\\odot}$) and high-mass (i.e., M $\\simeq$ 40\\,M$_{\\odot}$) SNII, respectively. In order to dig out the potential link between very metal-poor stars and isolated SNeII, Shigeyama \\& Tsujimoto (1998) also exploited the yields of core-collapse SN predicted by nucleosynthesis calculations (e.g., Woosley \\& Weaver 1995, Tsujimoto et al.~1995). They focused on the abundance patterns of C, Mg, Si and Ca. The yields of these elements being not (or at least less) affected by the mass-cut issue, they should therefore be known with a better accuracy than the yields of heavier elements such as the iron peak ones. They noticed the good agreement between the abundance patterns arising in the remnants of first generation SNeII, as predicted by nucleosynthesis calculations, and those inferred from the spectra of halo stars with [Fe/H] $\\lesssim -$2.5. As a result, Shigeyama \\& Tsujimoto (1998) suggested that very metal-poor stars actually formed out of interstellar gas which had been swept by isolated SNeII. They also note that the large observational scatter in the abundance ratios among these stars can be explained by the differences in SNII yields with the mass of the progenitor. It is worthy of a mention that such isolated SNeII are certainly not able to trigger the formation of GCs owing to the small amount of interstellar gas swept by the blast wave (Eq.~\\ref{single_SN}; also, not all the gas will be converted into stars). Should such low-mass halo clusters have managed to form however, they will have been quickly destroyed by the Galactic tidal fields, low-mass clusters being among the most vulnerable in this respect (e.g., Gnedin \\& Ostriker 1997). Therefore, this extension of the self-enrichement scenario to single SNII explosions does not contradict the absence of Galactic halo GCs with [Fe/H] $\\lesssim -$2.5. \\subsubsection{Hypernovae} In fact, even stars as massive as 40\\,M$_{\\odot}$ may trigger the formation of stars with a metallicity as low as $-4$, providing that they explode as {\\it hypernovae}, i.e., supernovae characterized by explosion energies of order $E_0 \\sim 10^{52} - 10^{53}$~ergs. This class of objects includes two categories depending on the progenitor mass, namely core-collapse hypernovae and pair-instability hypernovae. They both have been invoked as possible explanations to peculiar abundance patterns observed in some very metal-poor field stars. \\\\ {\\it Core-collapse hypernovae. } Recent observations suggest that at least some core-collapse supernovae explode with explosion energies ten to one hundred times higher than the energy released by a canonical SNII (e.g., Galama et al.~1998). They likely originate from relatively massive star (M$\\gtrsim 25$~M$_{\\odot}$). In contrast to pair-instability hypernovae (see below), the mass range of their progenitor is thus similar to the one of canonical SNeII. If hypernovae occured in the early stage of the Galactic evolution and induced star formation, their abundance pattern may still be observable in the atmospheres of some low-mass halo stars. As for the case of star formation triggered by single SNeII, we can derive an estimate of the metallicity of these halo stars using Eq.~\\ref{single_SN} and hypernova yields. Nakamura et al. (2001) have investigated in detail the nucleosynthesis of core-collapse hypernovae and compared it with the yields of canonical (i.e., $E_0 \\sim 10^{51}$\\,ergs) core-collapse SNeII with similar progenitor mass. Their Tables 2-5 show that SNeII and hypernovae with similar progenitor mass release much the same amount of metals. The resulting metallicity of stars formed in hypernova remnants are given in Table 1, following the same of line of reasoning as in the previous section. Hypernovae being much more energetic than SNeII, and the mass of interstellar gas swept by the blast wave being roughly proportional to the explosion energy (Eq.~\\ref{single_SN}), they will collect a larger amount of interstellar gas. Therefore, should core-collapse hypernovae be able to trigger the formation of new stars, these stars will be more metal-poor than the ones formed in the remnants of canonical SNeII. Thus, they may be well-suited to explain the formation of stars with a metallicity as low as [Fe/H] $\\simeq -4$ (see Table 1). \\begin{table} \\caption[]{Dependence of the metallicity [Fe/H] on the progenitor mass $m$ and explosion energy E$_0$ of supernovae (SN; E$_0 = 10^{51}$\\,ergs) and hypernovae (HN; E$_0 = 10 ~~ {\\rm or} ~~ 100 \\times 10^{51}$\\,ergs). Every mass is expressed in units of one solar mass. (1) A 12\\,M$_{\\odot}$ star explodes as a canonical SNII. According to Nakamura et al.~(2001), stars less massive than 25\\,M$_{\\odot}$ do not explode as core-collapse hypernovae and such cases are thus not considered in this Table. (2-3) More massive stars explode either as core-collapse supernovae or core-collapse hypernovae. (4) Very massive objects (i.e., M $>$ 100\\,M$_{\\odot}$) explode as pair-instability hypernovae. M$_{sw}$, the mass of interstellar gas collected by the explosion blast wave, depends on the explosion energy through Eq.~\\ref{single_SN}. m$_z$ is the mass of metals released by an exploding massive star with progenitor mass $m$. The metallicity [Fe/H] is derived assuming that the mass m$_z$ of metals is mixed with the mass M$_{sw}$ of collected interstellar gas} \\begin{center} \\begin{tabular}{ c c c c } \\hline E$_{0}/10^{51}$\\,ergs & 1 (SN) & 10 (HN) & 100 (HN) \\\\ \\hline M$_{sw}$ & 5.1 $\\times$ 10$^4$ & 5.1 $\\times$ 10$^5$ & 5.1 $\\times$ 10$^6$ \\\\ \\hline \\hline (1) (m,m$_z$)=(12,0.1)~ & & & \\\\ {\\rm [Fe/H]} $\\simeq$ & $-$4 & -- & -- \\\\ \\hline (2) (m,m$_z$)=(25,4)~~~ & & & \\\\ {\\rm [Fe/H]} $\\simeq$ & $-$2.4 & $-$3.4 & $-$4.4 \\\\ \\hline (3) (m,m$_z$)=(40,10)~~~ & & & \\\\ {\\rm [Fe/H]} $\\simeq$ & $-$2.1 & $-$3.1 & $-$4.1 \\\\ \\hline (4) (m,m$_z$)=(200,100) & & & \\\\ {\\rm [Fe/H]} $\\simeq$ & -- & $-$2.0 & $-$3.0 \\\\ \\hline \\end{tabular} \\end{center} \\end{table} The most significant feature of hypernova nucleosynthesis is their iron production, this one being larger than in SNeII by a factor 2 to 10 (Nakamura et al.~2001). This leads to small abundance ratios of $\\alpha$ elements over iron. Nakamura et al. (2001) thus suggested that the gas out of which the very metal-poor binary CS 22873-139 ([Fe/H] = $-$3.4) formed had been contaminated by the ejecta of an hypernova as this halo star shows almost solar [Mg/Fe] and [Ca/Fe] ratios. On the other hand, while stars with $-2.5 \\lesssim {\\rm [Fe/H]} \\lesssim 0$ show [Zn/Fe] $\\simeq$ 0 (e.g., Primas et al.~2000), this abundance ratio is steadily increasing towards [Zn/Fe] $\\sim$ 0.5 as the metallicity decreases, for stars more metal-poor than [Fe/H] $\\sim -2.5$ (e.g., Cayrel et al.~2003). Umeda \\& Nomoto (2002) notice that such a large [Zn/Fe] ratio arises naturally in their own hypernova model and thus conclude that core-collapse hypernovae are likely to have contributed to the early Galactic chemical evolution. \\\\ {\\it Pair instability hypernovae. } These very massive ($\\simeq$ 140-260\\,M$_{\\odot}$) objects exploding with $E \\gtrsim 10^{52}$\\,ergs, they are also called hypernovae. Their name (i.e., ``pair instability'') refers to the electron-positron pair instability process encountered during the central oxygen-burning stages (see Umeda \\& Nomoto 2002, their appendix for a detailed description). The main feature of these very massive stars is their ability to ``pass carbon and oxygen from the helium-burning core through the hydrogen-burning shell, in such a way that it is CNO processed to nitrogen before entering the hydrogen envelope'' (Carr, Bond \\& Arnett 1984). These stars thus produce large supersolar values of N/Fe, an effect not predicted by models of Galactic chemical enrichment based on stars less massive than 100\\,M$_{\\odot}$. Following their discovery of CS 22949-037, a very metal-poor star ([Fe/H] = $-$3.8) showing extreme nitrogen enhancement ([N/Fe] = 2.3), Norris et al. (2002) suggested that this star may have been formed out of gas polluted and compressed by such a very massive hypernova. On the other hand, in marked contrast with core-collapse hypernova yields, Umeda \\& Nomoto (2002) quote that pair-instability hypernovae are unlikely to produce [Zn/Fe] ratios as large as in very metal-poor stars where [Zn/Fe] ranges from solar up to $\\simeq$0.5 (Cayrel et al.~2003). Actually, the abundance ratio derived from their hypernova yields is of order [Zn/Fe]=$-$1.5, that is, 1 to 2 dex smaller than in very metal-poor stars. Additionally, Heger \\& Woosley (2002) note the absence of the r-process in these massive objects, which conflicts with observations showing appreciable amounts of r-process elements in very metal-poor stars (e.g., Burris et al.~2000). Based on these two arguments, i.e., the low abundance ratio [Zn/Fe] as well as the absence of r-process elements in pair instability hypernova ejecta, Umeda \\& Nomoto (2002) and Heger \\& Woosley (2002) conclude that the abundances of very metal-poor stars cannot be ascribed to pair instability hypernovae only and must include the contribution of an additional nucleosynthetic component, namely those lower mass stars that make ``regular'' supernovae. We also note that pair-instability hypernovae release fairly large amounts of metals in the interstellar medium, of order 100~M$_{\\odot}$ (Umeda \\& Nomoto 2002, their Tables 15-18). In fact, these stars disrupt completely when exploding, leaving no compact remnant after the explosion (i.e., no issue of ``mass-cut'' or ``fall-back'', see Heger \\& Woosley 2002). The mixing of so large an amount of metals with a mass of primordial gas of $\\simeq 5 \\times 10^5 {\\rm M}_{\\odot}$ (E$_{0} = 10 \\times 10^{51}$\\,ergs) or $\\simeq 5 \\times 10^6 {\\rm M}_{\\odot}$ (E$_{0} = 100 \\times 10^{51}$\\,ergs) will lead to metallicities of order [Fe/H] $\\simeq -2$ or [Fe/H] $\\simeq -3$, respectively. Therefore, the metallicity of the most metal-poor stars in the Galactic halo cannot be explained by pair-instability hypernovae. \\\\ \\subsubsection{Single short-acting source of energy and triggered star formation} The issue of whether isolated exploding massive or very massive stars can stimulate the formation of new stars in the layers of interstellar gas they have swept is not explicitely addressed in the papers mentioned in Sect.~3.2.1 and 3.2.2 (e.g., Shigeyama \\& Tsujimoto 1998, Argast et al.~2000, Norris et al.~2002). It is merely assumed that such stimulated star formation actually took place during the early Galactic stages. The model presented in this paper (see sect.~2) is not the most appropriate to address this issue quantitatively as it deals with a continuous input energy and not with a single short-acting source of energy. Actually, supershells created in connection to single explosion and those created around a cluster of massive stars do not show the same expansion rate with time (see below). We have seen in Sect.~\\ref{sec:stimSF} how the radius and the velocity of the shell affect its ability to collapse through the shell surface density and the stretching of the forming perturbation, respectively. Hence, to solve the issue of whether a single supernova/hypernova triggers the formation of new stars requires to derive a new temporal evolution of the shell radius, in the PGCC and in the protogalactic background, adequate to single explosion. This is beyond the scope of the present paper whose main goal is to address the formation of stars with $-2.5 \\lesssim {\\rm [Fe/H]} \\lesssim -1$ (i.e., the metallicity range of halo GCs). Some qualitative inferences can however be made. Efremov, Ehlerova \\& Palous (1999) compared the propagation in an homogeneous medium of both kinds of supershells. In case of an abrupt energy input, the early expansion of the supershell proceeds at a higher velocity than in the case of a continuous input of the same energy amount. Accordingly, in the former case, the shell of swept gas is more strongly stretched and this hampers the formation of an initial perturbation. After a few millions years however, the shell created in connection to an abrupt energy input slows down and shows both a radius and a velocity lower than in case of a continuous energy input. In contrast to the initial shell expansion, this second stage favours the transverse collapse through a larger shell surface density and a weaker stretching of the shell perturbed regions. Whether the collapse will occur is actually not certain. The source of the energy input being point-like (i.e., its size is much smaller than the one of a cluster of massive stars), the initial amplitude of the transverse motions within the shell will be lowered accordingly, thus reducing the collapse. Comparing SNeII and hypernovae, hypernovae sweep a larger amount of gas thanks to their larger energy input (Eq.~\\ref{single_SN}). However, this also leads to a larger shell radius. Both effects acting on the shell surface density in opposite ways, only detail computations will tell us which class of objects is the most efficient in forming new stars. A $E_0 \\sim 2 \\times 10^{52}$\\,ergs hypernova being able to collect amount of gas as large as 10$^6$\\,M$_{\\odot}$ (Eq.~$\\ref{single_SN}$), one might think that pair-instability hypernovae could be related to the formation of GCs, at least the most metal-poor ones as such hypernovae can chemically enrich the initially pristine gas up to [Fe/H] $\\simeq$ -2 (see Table 1). However, the [Zn/Fe] abundance ratio predicted for hypernova ejecta (i.e., [Zn/Fe] $<-1.5$, Umeda \\& Nomoto 2002) shows a sharp discrepancy with the roughly solar value observed in halo field stars with [Fe/H] $\\simeq -2$ (Primas et al.~2000) as well as in NGC6397, a halo GC with the same metallicity (Thevenin et al.~2001). Therefore, the abundance patterns of stars in this metallicity range cannot be ascribed to pair-instability hypernovae. Obviously, it is not straightforward to conclude whether single massive star explosions are able to trigger the formation of new stars. More computations are required to answer a question which is certainly worthy of further investigation owing to its potential link with very metal-poor stars. \\subsubsection{External pollution onto PopIII stars} Shigeyama, Tsujimoto \\& Yoshii (2003) have recently suggested that some very metal-deficient stars may have been born as metal-free (PopIII) stars whose external layers were afterwards polluted by the accretion of chemically enriched interstellar gas. The metallicity achieved by the external layers will depend on the metallicity and mass of the accreted gas, as well as on the stellar mass fraction with which the accreted gas has been mixed. As a consequence, the chemical enrichment of stellar superficial layers will show up more markedly in main sequence stars than in red giants, the convective envelope (i.e. the mixing zone) being almost two orders of magnitude larger in halo giants (about half the star mass) than in non-evolved metal-poor stars (about one per cent of the star mass). Shigeyama et al.~(2003) have shown that the extremely low metallicity ([Fe/H]$\\simeq -$5.3) of the giant HE0107-5240 (Christlieb et al.~2002) may originate from such a mechanism, that is, an external pollution while the star was on the main sequence followed by the dilution of the accreted material as the star started ascending the red giant branch. Owing to their thin convective envelopes, polluted Pop III stars still on the main sequence would exhibit larger metallicities, i.e. of order [Fe/H]$\\simeq -$3, although still lower than the most metal-poor GCs. \\\\ \\subsubsection{Star formation in dwarf galaxies} Within the frame of a hierarchical model for halo formation, Cote et al.~(1999) proposed that the excess of very metal-poor field stars with respect to GCs was formed in the most metal-deficient dwarf galaxies trapped within the Galactic potential well. These galaxies are not expected to contribute to the GC population. Indeed, due to the luminosity-metallicity correlation observed among dwarf galaxies (Gilmore 2000, Mateo 2000), the most metal-poor dwarfs are also expected to be the least luminous. Dwarf galaxies exhibit a luminosity threshold for hosting a GC system. For instance, the Fornax dwarf spheroidal galaxy is the faintest ($M_v \\simeq -12.3$, Harris 1991) galaxy known to harbour its own GC system. Therefore, the dimmest dwarf galaxies, which are also the most metal-poor, may have contributed to the halo field and not to the halo GC system. This paper has presented the results of simulations dedicated to the transverse collapse of shells of gas resulting from the sweeping of gaseous GC progenitors by SNeII. In these simulations, the growth of an initial perturbation in the shell surface density is followed by solving the linear perturbed equations of continuity and motion for transverse flows in a spherical shell of gas. Such a collapse depends on several parameters, namely the number of SNeII, the background presure, the sound speed of the shell gas and the initial conditions of the perturbation (i.e. the number of clumps, the initial perturbed surface density and velocity and the corresponding phase difference). All these parameters have been discussed in turn. The results show that the pressure $P_h$ of the hot protogalactic background ($P_h \\sim 10^{-10}$\\,dyne.cm$^{-2}$, Fall \\& Rees 1985, Murray \\& Lin 1992) and the numbers $N$ of SNeII allowed by the disruption criterion (i.e. smaller than 200, Paper I) can indeed lead to a successful shell transverse collapse, and thereby to the formation of new stars, assuming some reasonable initial conditions for the perturbation (see Fig.~8). The metallicities achieved in the shells able to collapse agrees with the metallicity range of Galactic halo GCs, namely, $-2.5 \\lesssim {\\rm [Fe/H]} \\lesssim -1$. Furthermore, while $N$ and $P_h$ determine the metallicity achieved through self-enrichment, they also control the probability of triggered star formation and the ability of these second generation stars to form a bound GC. Such a property is the most interesting since it opens the way to the understanding of the halo metallicity distribution functions, for both stars and clusters." }, "0403/astro-ph0403649.txt": { "abstract": "{ Magnetic field behaviour in a spherically-symmetric accretion flow for parameters typical of single black holes in the Galaxy is discussed. It is shown that in the majority of Galaxy volume, accretion onto single stellar-mass black holes will be spherical and have a low accretion rate ($10^{-6} - 10^{-9}$ of the Eddington rate). An analysis of plasma internal energy growth during the infall is performed. Adiabatic heating of collisionless accretion flow due to magnetic adiabatic invariant conservation is $25\\%$ more efficient than in the standard non-magnetized gas case. It is shown that magnetic field line reconnections in discrete current sheets lead to significant nonthermal electron component formation. In a framework of quasi-diffusion acceleration, the \"energy-radius\" electron distribution is computed and the function describing the shape of synchrotron radiation spectrum is constructed. It is shown that nonthermal electron emission leads to formation of a hard (UV, X-ray, up to gamma), highly variable spectral component in addition to the standard synchrotron optical component first derived by Shvartsman generated by thermal electrons in the magnetic field of accretion flow. For typical interstellar medium parameters, a black hole at 100 pc distance will be a 16-25$^{\\rm m}$ optical source coinciding with the highly variable bright X-ray counterpart, while the variable component of optical emission will be about 18-27$^{\\rm m}$. The typical time scale of the variability is $10^{-4}$ sec, with relative flare amplitudes of 0.2-6\\% in various spectral bands. Possible applications of these results to the problem of search for single black holes are discussed. ", "introduction": "\\label{sec_introduction} Even though more than 60 years have passed since the theoretical prediction of black holes as an astrophysical objects (Oppenheimer \\& Snyder \\cite{oppenheimer}) in some sense they have not been discovered yet. To identify an object as a black hole, one needs to show that its mass exceeds $3M_{\\odot}$, its size is close to $r_g = 2G M/c^2$ and it has an event horizon instead of a normal surface -- the distinguishing property of black holes which separates them from massive compact objects of finite size in some theories of gravity (Will \\cite{will}). However, only the two former criteria are used now for selection of black hole candidates of two types: a) with masses of 5-18 $M_{\\odot}$, in X-ray binaries (see, for example, Greiner et al. \\cite{greiner}); and b) supermassive black holes in galaxy nuclei with masses of $10^6 - 10^{10} M_{\\odot}$ (Shields \\cite{shields_1999}). Existence of the event horizon in such objects is usually implied by the absence of periodic pulsations of the X-ray emission from strong regular magnetic fields (the black hole \"no-hair\" theorem) and I type X-ray flares due to thermonuclear bursts of the accreted matter on the surface of the neutron star. At the same time, typical masses of X-ray pulsars and bursters are close to the typical neutron star value of 1.4 $M_{\\odot}$ while black hole candidates, missing pulsations and X-ray flares, have masses of 5-18 $M_{\\odot}$ (Miller et al. \\cite{miller_1998}). The absence of an event horizon in low-mass objects is not a proof of its existence in higher-mass ones. High accretion rates in X-ray binaries and active galactic nuclei result in the screening of regions close to the event horizon, and the most luminous parts of accretion flow are situated at distances of $10 - 100 r_g$ (Chakrabarti \\cite{chakrabarti}; Cherepashchuk \\cite{cherepashchuk}) where general relativity effects are negligible. There is a very effective way to get information about the innermost parts of accretion disks in X-ray binaries as well as AGNs -- the investigation of the broad (and sharp) iron $K_{\\alpha}$ fluorescent emission line (see review by Reynolds \\& Nowak \\cite{reynolds_2003}). Its intensity and shape depend on the accreted plasma distribution and behaviour until the last stable orbit (0.62 $r_g$ for a extremely spinning Kerr black hole and 3 $r_g$ for a Schwarzschild one) (Miller et al. \\cite{miller_2004}; Miniutti et al. \\cite{miniutti}). However since the photons generated at the different distances from the horizon are mixed in the line profile it is not possible to extract the manifestations of gravitational fields close to the horizon only. This may be possible by study of variability of the iron line (Reynolds et al. \\cite{reynolds_2004}). At the same time, single stellar-mass black holes, which accrete interstellar medium of low density ($10^{-2} - 1 $cm$^{-3}$), are the ideal case for detection and study of the event horizon. Shvartsman (\\cite{shvartsman_1971}) first demonstrated that an emitting halo of accreted matter forms around such objects and generates optical featureless emission. The majority of such emission comes from the regions near the horizon at $(3 -- 5) r_g$. Spherical accretion onto the single stellar-mass black holes has been studied in detail in the works of several authors (Bisnovatyi-Kogan \\& Ruzmaikin \\cite{bisnovatyi_1974}; Meszaros \\cite{meszaros}; Ipser \\& Price \\cite{ipser_1977, ipser_1982}) and the main conclusions of Shvartsman have been confirmed. The most striking property of the accretion flow onto the single black hole is its inhomogeneity -- the clots of plasma act as a probe testing the space-time properties near the horizon. The characteristic timescale of emission variability is $\\tau_v\\sim r_g/c\\sim 10^{-4} - 10^{-5}$ sec and such short stochastic variability may be considered as a distinctive property of black hole as the smallest possible physical object with a given mass. Its parameters -- spectra, energy distribution and light curves -- carry important information on space-time properties of the horizon (Beskin \\& Shvartsman \\cite{beskin_1976}). The general observational appearance of a single stellar-mass black hole at typical interstellar medium densities is the same as other optical objects without spectral lines -- DC-dwarfs and ROCOSes (Radio Objects with Continuous Optical Spectra, a subclass of blazars) (Beskin \\& Mitronova \\cite{beskin_1991}, Pustilnik \\cite{pustilnik_1977}, Shvartsman \\cite{shvartsman_1977}; Beskin et al. \\cite{beskin_2000}). The suggestion that isolated BHs can be among them is the basis of the observational programme of search for isolated stellar-mass black holes -- MANIA (Multichannel Analysis of Nanosecond Intensity Alterations). It uses photometric observations of candidate objects with high time resolution, special hardware and data analysis methods (Shvartsman \\cite{shvartsman_1977}; Beskin et al. \\cite{beskin_1997}). In observations using the 6-meter telescope of the Special Astrophysical Observatory of 40 DC-dwarfs and ROCOSes, only upper limits for variability levels of 20\\% -- 5\\% on the timescales of $10^{-6}$ -- 10 sec, respectively, were obtained, i.e. BHs were not detected (Shvartsman et al. \\cite{shvartsman_1989a}, \\cite{shvartsman_1989b}, Beskin et al. \\cite{beskin_2000}). Recently, some evidences appeared that single stellar-mass black holes may be found among the stationary unidentified gamma-ray sources (Gehrels et al. 2002), gravitational lenses causing long-lasting MACHO events (Bennett et al. 2001) and white dwarf -- black hole binaries detected by means of self-microlensing flashes (Beskin \\& Tuntsov \\cite{beskin_2002b}). In the last case, the mass transfer from the white dwarf is absent and a black hole behaves as a single one. In the present work we study spherical accretion onto a single stellar-mass black hole at low accretion rates of $10^{8} - 10^{13}$ g/s. This corresponds to the range of interstellar medium densities of $0.002 - 0.1$cm$^{-3}$ and a 10 $M_{\\odot}$ object moving with a velocity of 20--40 km/s (Bondi \\& Hoyle \\cite{bondi_1944}). This is true for about 90\\% of the Galaxy volume (McKee \\& Ostriker \\cite{mckee}). Spherical accretion with equipartition of energies, i.e. with roughly equal densities of the magnetic and kinetic energies of plasma has been considered in many papers (Shvartsman \\cite{shvartsman_1971}; Bisnovatyi-Kogan \\& Ruzmaikin \\cite{bisnovatyi_1974}; Kowalenko \\& Melia \\cite{kowalenko}; Ipser \\& Price \\cite{ipser_1977,ipser_1982}). The uniqueness of our approach is in taking into account the significantly non-thermal nature of electron energy distribution function (its synchrotron emission determines the appearance of a black hole). It may be roughly considered as a superposition of two components (this approach is known as \"hybrid plasma\", see Coppi (\\cite{coppi}) and references therein) -- thermal electrons and accelerated electron beams, formed in current sheets where magnetic energy is dissipated in a way similar to solar flares (Pustilnik \\cite{pustilnik_1978,pustilnik_1997}). The latter process supports the equipartition of energies. As a result, the emission of the accretion flow consists of a quasi-stationary \"thermal\" part with a wide-band spectrum from infrared to ultraviolet, and a highly variable flaring nonthermal component. Each such flare is generated due to the motion of the accelerated electron beam in the magnetic field. Its light curve carries information on the magnetic and gravitational field structure near the black hole horizon. Nonthermal luminosity reaches several percents of the total luminosity and may even exceeds it at low rates, while its spectrum covers spectral bands from optical to hard X-ray. This result leads to possible modifications of the search strategy. In \\S 2, the main characteristics of accretion onto single black holes in the Galaxy are discussed. In \\S 3, the electron distribution function in phase space is built, in \\S 4, the thermal and nonthermal component luminosities are determined, and in \\S 5 the shape of its spectra is studied. In \\S 6, the temporal behaviour of single electron beam emission is studied and some conclusions on the variability of accretion flow are made. In \\S 7, the main results of this work are summarized, and in \\S 9, possible directions of future work are discussed. ", "conclusions": "During recent years the number of works dealing with single stellar-mass black holes has significantly increased. Some are purely theoretical (Punsly \\cite{punsly_1998a,punsly_1998b}; Gruzinov \\& Quataert \\cite{gruzinov}; Abramowicz et al. \\cite{abramowicz}) and other provide discussions of their observational detection (Heckler \\& Colb \\cite{heckler}; Fujita et al. \\cite{fujita}; Beskin et al. \\cite{beskin_2000}; Agol \\& Kamionkowski \\cite{agol_2002b}; Chisholm et al. \\cite{chisholm}). The importance of experiments in strong gravitational fields has been recently noted by Damour (\\cite{damour}). Kramer et al. (\\cite{kramer}) discussed the new possibilities of black holes metric study by the investigation of radio pulsar -- BH binary systems with new generation of radio telescopes. In this work we tried to concretize physical properties of plasma accreted onto the black hole within the classical paradigm of equipartition of Shvartsman (\\cite{shvartsman_1971}). Assuming the discrete nature of the magnetic energy dissipation processes in current sheets allows us to clarify the shape of the synchrotron spectrum of the accretion flow. A hard highly non-stationary nonthermal spectral component appears as an emission of accelerated particles. The beams accelerated in the current sheets can generate very short flares, providing information about the neighborhood of the event horizon (Fig.\\ref{fig_flare_lightcurve}). On the other hand it is clear from Fig.\\ref{fig_luminosity_radial} that at low accretion rates a significant amount of thermal synchrotron radiation is generated inside $3r_g$ -- this means that the behaviour of this component will reflect the properties of space-time in strong gravitational fields too. It is clear that the search for a black hole strategy may be modified in accordance with such results. Optical high time resolution studies of X-ray sources may be very important. Single black holes may be contained inside the known stationary gamma sources (Gehrels et al. \\cite{gehrels}) as well as objects causing long microlensing events (Paczynski \\cite{paczynski}). Thus it is very important to look for X-ray emission as well as for fast optical variability of these objects. Sample observations of the longest microlensing event MACHO 1999-BLG-22 (Bennett et al. \\cite{bennett_2001}), a stellar-mass black hole candidate, have been performed at the Special Astrophysical Observatory of RAS in the framework of the MANIA experiment in 2003-2004 (Beskin et al. \\cite{beskin_2005}). The best evidence will be provided by the synchronous high time resolution observations in optical and X-ray ranges. Detection of the event horizon signatures cannot result from statistical studies. A detailed study of each object is needed to detect its specific appearance." }, "0403/astro-ph0403528_arXiv.txt": { "abstract": "Annihilating dark matter particles produce roughly as much power in electrons and positrons as in gamma ray photons. The charged particles lose essentially all of their energy to inverse Compton and synchrotron processes in the galactic environment. We discuss the diffuse signature of dark matter annihilations in satellites of the Milky Way (which may be optically dark with few or no stars), providing a tail of emission trailing the satellite in its orbit. Inverse Compton processes provide X-rays and gamma rays, and synchrotron emission at radio wavelengths might be seen. We discuss the possibility of detecting these signals with current and future observations, in particular EGRET and GLAST for the gamma rays. ", "introduction": "It is almost universally accepted that most of the matter in the universe is non-baryonic. This dark matter is the chief constituent of gravitationally bound objects from dwarf galaxy scales and larger. Identifying the nature of dark matter is one of the most important problems in astrophysics, cosmology, and particle physics. Perhaps the best motivated candidate for cold dark matter is the lightest of the so--called neutralinos arising in supersymmetric extensions to the standard model \\cite{jkg96}. These are the spin-1/2 Majorana fermion counterparts of the neutral gauge and Higgs bosons, and are expected to have masses at the weak scale (of order 100 GeV). This scale is intriguing as the relic density of a stable particle with this mass and corresponding cross section turns out to be of order the critical density, as we observe the matter density to be today. Any stable particles with weak scale masses could thus naturally account for the dark matter. In this paper, we will focus on supersymmetry, but our conclusions are fairly generic to dark matter candidates at the TeV scale. We outline a new signature of annihilating dark matter in satellites of the Milky Way galaxy. Prospects for detecting high energy photons as dark matter annihilation products, primarily from the $\\pi^0$ decays that are generic to hadronization processes, have been discussed for many years (for a sample see Ref.~\\cite{gammas}). Necessarily coming with these photons are high energy electrons and positrons from the analogous $\\pi^\\pm$ decay chains. Charged particles suffer complicated motions in the galactic magnetic field, and furthermore they lose energy to synchrotron and inverse Compton processes. Searching for the synchrotron emission from the galactic center \\cite{syncGC} and from galactic satellites \\cite{blasi} has been discussed previously, though neglecting the diffusion of the charged particles. We will show that the inverse Compton emission, extended over a large area from the charged particle annihilation products may be observable for some models of particle dark matter and of galactic satellites. ", "conclusions": "We have calculated the spectrum and spatial extent of diffuse emission from the charged particle products of dark matter annihilations. In addition to the synchrotron emission discussed previously, we have studied the inverse Compton radiation, primarily on starlight photons. We have focused on galactic satellites that are currently within the diffusion zone, namely within a few kpc of the stellar disk. For satellites moving with typical galactic halo velocities of 300 km s$^{-1}$, the crossing time of the diffusion zone is of the same order as the diffusion time, thus an inherently time-dependent treatment is required. For annihilation sources, e.g.\\ galactic satellites at typical distances of 10 kpc, the diffuse emission in both inverse Compton and synchrotron extends over roughly 300 square degrees. We have shown that at least in terms of the number of photons, the diffuse inverse Compton emission might be detectable by GLAST, assuming bright enough annihilation sources. The spatial extent of the emission makes its detection problematic of course. GLAST will certainly detect a significant number of point sources in a region of this size. In a future work we will study in detail the feasibility of separating these signals. As mentioned previously, these results are fairly generic, and do not depend strongly on the particle physics model. As we are concerned with the electrons and positrons, we do not even require that the dark matter have hadronic interactions. Leptonically interacting dark matter \\cite{krauss,baltzbergstrom} would still provide photons and electrons, albeit by different processes and with different spectra. Such photon sources would be even harder to reconcile with the EGRET point sources, but annihilation sources below the EGRET detection limit may be detectable by GLAST in any case." }, "0403/astro-ph0403113.txt": { "abstract": "We have surveyed $\\sim 400 $ known large-amplitude variables within 15$'$ of the galactic center in the SiO $J=1$--0 $v=$ 1 and 2 maser lines at 43 GHz, % resulting in 179 detections. SiO lines were also detected from 16 other resulting in 180 detections. SiO lines were also detected from 16 other sources, which are located within 20$''$ (the telescope half beamwidth) of the program objects. The detection rate of 48 percent is comparable to that obtained in Bulge IRAS source surveys. Among the SiO detections, five stars have radial velocities greater than 200 km s$^{-1}$. The SiO detection rate increases steeply with the period of light variation, particularly for stars with $P>500$ d, where it exceeds 80\\%. We found that, at a given period, the SiO detection rate is approximately three times that for OH. These facts suggest that the large-amplitude variables in the Nuclear Disk region are AGB stars similar in their overall properties to the inner and outer Bulge IRAS/SiO sources. From the set of radial velocity data, the mass distribution within 30 pc of the galactic center is derived by a new method which is based on the collisionless Boltzmann equation integrated along the line of sight. The mass within 30 pc is about $6.4 [\\pm 0.7] \\times 10^7 $ M$_{\\odot}$ and the mass of the central black hole is $2.7 [\\pm 1.3] \\times 10^6 $ M$_{\\odot}$. Consideration of the line-of-sight velocity of each star and its potential energy leads to the conclusion that the five high-velocity stars come from galactocentric distances as high as 300 pc. The high-velocity subsample of stars with negative radial velocities exhibits a tendency to have brighter $K$ magnitudes than the subsample of stars with positive velocities. The origin of these high-velocity stars is discussed. ", "introduction": "Radial-velocity data concerning stellar maser sources are useful for studying the dynamical behavior of the central part of the Galaxy (\\cite{lin92a}; \\cite{izu95}; \\cite{sjo98}; \\cite{deg00}). At visible wavelengths, they are difficult to obtain in the galactic center region because of interstellar extinction. Instead, most information comes from radio or near-infrared observations [for example, \\citet{sel87}]. In particular, observations of SiO and OH masers give radial velocities of stars accurate to within a few km s$^{-1}$. The masers arise in the circumstellar envelopes of mass-losing stars on the Asymptotic Giant Branch (AGB), which are intrinsically bright in the near- and mid-infrared regions, and which can potentially be identified at these wavelengths. Large numbers of candidate stars suitable for pointed maser surveys toward the nuclear disk have been discovered in the near-infrared $K$ band by making use of their characteristic large-amplitude variability (\\cite{gla01}). The dynamical behavior of the central region of the Galaxy has attracted much attention, especially in relation to the central black hole (for example, \\cite{mor96}). Proper motions of stars have been measured in the near-infrared $K$ band near the black hole (\\cite{gen00}; \\cite{ghe00}), and proper motions of SiO maser stars have also been measured within the central 15$''$ (\\cite{rei03}). These were used to find the position of Sgr A* in the $K$-band images (\\cite{men97}). This paper concentrates on the dynamics of stars located towards the outer part of the central star cluster around the black hole, i.e., at about 2--30 pc distance, where the gravitational force of the black hole ceases to influence the stellar motions, and the stellar system is nearly self-gravitating. The dynamical (rotational) time scale in this region is a few $\\times 10^6$ y, while the ages of the AGB stars are $10^7$ -- $10^9$ y. Therefore, these stars are considered to be dynamically well relaxed (e.g., \\cite{hoz00}). Since the bar-like structure of the Galactic bulge was discovered (\\cite{bli91}; \\cite{nak91}; \\cite{dwe95}), it has been recognized that non-circular motions must be taken into account when interpreting observational data such as the CO gas distribution in the central nuclear disk (\\cite{bin91}; \\cite{wei99}). Because double bars and nuclear rings have been proposed as efficient mechanisms for feeding gas into the centers of galaxies (\\cite{shl89}), it has become additionally important to look for signs of non-circularity in the motions of gas and stars. In this paper, we report on the results of an SiO maser survey of Large Amplitude Variables (Miras or semiregulars; abbreviated as LAV hereafter) in a $24' \\times 24'$ area of the galactic center (\\cite{gla01}), whose amplitudes and periods are known (\\cite{woo98}; \\cite{gla01}). Because these stars are located at approximately the same distance (about 8 kpc) from the Sun, they constitute an ideal sample for studying the statistical characteristics of AGB stars and their detectability in the maser lines. In addition, surveying these sources gives accurate radial velocities, and provides basic data for investigating the kinematics of the galactic nuclear disk. Although the mass within this region has been obtained previously by various methods [for example, \\citet{lin92a}], none of them [except \\citet{sah96}] are fully valid for treating the problem; the results obtained are likely to be in error by a factor of a few between the radii of 2 and 30pc. In this paper, we analyze the new SiO radial-velocity data set using the Boltzmann equation integrated along the line of sight. We also consider the origin of the high-velocity stars seen toward the galactic center. %The position of the Sgr A*, which is considered to be at the dynamical center of the Galaxy, %is at ($RA$, $Dec$, epoch)= (17h45m40.045s, $-29^{\\circ}00'27.9''$, J2000). Because Sgr A* is %shifted to the origin of galactic coordinates by about $0.05^{\\circ}$, it is more convenient %to use the coordinates in which Sgr A* is located at the origin. For this purpose, %we use the symbol, $l^* (=l+0.056^{\\circ})$, and $b* (=b+0.046^{\\circ}$ in this paper. %\\newpage ", "conclusions": "We have surveyed $\\sim 400$ large-amplitude variables within a $24'\\times 24'$ square about the galactic center, and obtained 180 detections (with additional 16 detections other than LAVs) in the SiO maser lines. The SiO detection rate of $\\sim$48\\% is comparable to that in previous SiO surveys of color-selected Bulge IRAS sources. The SiO detection rate increases with the period of light variation, and is well correlated with the OH detection rate. The longitude-velocity diagram of the SiO sources has been revealed to be quite similar to the OH $l$--$v$ diagram. These facts suggest that the large-amplitude variables in the galactic nuclear disk are mass-losing stars in the AGB phase, quite similar to the IRAS sources in the inner Galactic bulge. We also analyzed the SiO radial velocity data and obtained the mass distribution of the galactic center area. The mass of the central black hole that we have deduced, $2.7 (\\pm1.3) \\times 10^6 M_{\\odot}$, and the mass within 30 pc, $6.5 (\\pm 0.7) \\times 10^7 M_{\\odot}$, are more accurate than the previous estimates for these quantities. From analysis of the projected mass vs radius diagram, we found a tendency among the high-velocity sources that the subset with negative line-of-sight velocity is systematically brighter than the subset with positive line-of-sight velocity. This results from the fact that the the subsample with negative line-of-sight velocity is in front of the galactic center and the subsample with positive velocity is behind it. This tendency, which also applies to the Bulge SiO maser sources, strongly suggest the presence of streaming motion in the present nuclear-disk LAV sample. The authors thank Dr. A. Winnberg for the useful comments. One of authors (I.S.G.) thanks the National Astronomical Observatory for providing him with a visiting fellowship for this work. This research was made use of the SIMBAD database operated at CDS, Strasbourg, France. It was partly supported by Scientific Research Grant (C2) 12640243 of the Japan Society for Promotion of Sciences. %%%%%%%%%% Appendix %%%%%%%%%%%%%" }, "0403/astro-ph0403658_arXiv.txt": { "abstract": "We analyze the statistical properties and dynamical implications of galaxy distributions in phase space for samples selected from the 2MASS Extended Source Catalog. The galaxy distribution is decomposed into modes $\\delta({\\bf k, x})$ which describe the number density perturbations of galaxies in phase space cell given by scale band $\\bf k$ to ${\\bf k}+\\Delta {\\bf k}$ and spatial range $\\bf x$ to ${\\bf x}+\\Delta {\\bf x}$. In the nonlinear regime, $\\delta({\\bf k, x})$ is highly non-Gaussian. We find, however, that the correlations between $\\delta({\\bf k, x})$ and $\\delta({\\bf k', x'})$ are always very weak if the spatial ranges (${\\bf x}$, ${\\bf x}+\\Delta {\\bf x}$) and (${\\bf x'}$, ${\\bf x'}+\\Delta {\\bf x'}$) don't overlap. This feature is due to the fact that the spatial locality of the initial perturbations is memorized during hierarchical clustering. The highly spatial locality of the 2MASS galaxy correlations is a strong evidence for the initial perturbations of the cosmic mass field being spatially localized, and therefore, consistent with a Gaussian initial perturbations on scales as small as about 0.1 h$^{-1}$ Mpc. Moreover, the 2MASS galaxy spatial locality indicates that the relationship between density perturbations of galaxies and the underlying dark matter should be localized in phase space. That is, for a structure consisting of perturbations on scales from $k$ to $ k+\\Delta { k}$, the nonlocal range in the relation between galaxies and dark matter should {\\it not} be larger than $|{\\Delta {\\bf x}}|=2\\pi/|\\Delta {\\bf k}|$. The stochasticity and nonlocality of the bias relation between galaxies and dark matter fields should be no more than the allowed range given by the uncertainty relation $|{\\Delta {\\bf x}|| \\Delta{\\bf k}}|=2\\pi$. ", "introduction": "The large scale structure of the universe developed from initial mass density and velocity fluctuations through gravitational instability. Much of the information of the initial perturbations is ``forgotten\" during the gravitational nonlinear evolution. Yet, some features of the initial perturbations are imprinted in the cosmic mass and velocity fields at present. Measuring these memorized features is crucial in studying the initial states of the universe. A well known example is the self-similarity of gravitational clustering, which imprints the initial power spectrum index and is detectable with the scaling behavior of correlation functions of the present mass field (e.g. Peebles 1980). In this paper, we study the spatial locality of the perturbed mass field in phase space, which is also a feature to be memorized in mass field today. The physics of spatial locality of correlation function in phase space can be illustrated with a Gaussian mass field $\\rho({\\bf x})$ described by \\begin{equation} \\langle\\hat{\\delta}({\\bf k}) \\hat{\\delta}({\\bf -k'})\\rangle = P(k)\\delta^K_{\\bf k,k'}, \\end{equation} where $\\hat{\\delta}({\\bf k})$ is the Fourier counterpart of the density contrast $\\delta({\\bf x})=[\\rho({\\bf x})-\\overline\\rho]/\\overline{\\rho}$, $P(k)$ the power spectrum of the mass field, and $\\delta^K$ the Kronecker delta function. Eq.(1) says that the Fourier modes with different wavevector ${\\bf k}$ are uncorrelated, or the correlation is localized in $k$-space. This is because the phase of the Fourier modes $\\delta({\\bf k})$ is random. On the other hand, the correlation function of density perturbations in physical ($x$) space generally is non-local. The two-point correlation function $\\langle \\delta({\\bf x})\\delta({\\bf x'})\\rangle$ has non-zero correlation length when the Fourier power spectrum $P(k)$ of eq.(1) is $k$-dependent. In a phase-space description, the mass field is decomposed into modes $\\delta({\\bf k, x})$, the perturbations in the wavevector(scale) from ${\\bf k}$ to ${\\bf k} + \\Delta {\\bf k}$ and physical range ${\\bf x}$ to ${\\bf x}+\\Delta {\\bf x}$. The volume of the phase space cell referring to a mode is given by the uncertainty relation $|\\Delta {\\bf x}||\\Delta {\\bf k}| =2\\pi$. The correlation function of a Gaussian field generally is localized regardless whether the Fourier power spectrum is colored. That is, \\begin{equation} \\langle \\delta({\\bf k, x})\\delta({\\bf k', x'})\\rangle \\propto \\delta^K_{\\bf x,x'}\\delta^K_{\\bf k,k'}. \\end{equation} The reason for eq.(2) is straightforward. First, the perturbations $\\delta({\\bf k, x})$ and $\\delta({\\bf k', x'})$ are, respectively, given by linear superposition of the Fourier modes in different wavebands (${\\bf k}$, ${\\bf k}+ \\Delta {\\bf k}$) and (${\\bf k'}$, ${\\bf k'}+ \\Delta {\\bf k'}$). For a Gaussian field, the Fourier modes in different wave band are uncorrelated in general [eq.(1)]. This gives rise to the factor of $\\delta^K_{\\bf k, k'}$ of eq.(2). Second, the phases of the Fourier modes of Gaussian field are random. For a superposition of the Fourier modes in the band (${\\bf k}$, ${\\bf k} +\\Delta {\\bf k}$) with random phase, the coherent length of the phase of $\\delta({\\bf k,x})$ is not larger than $|\\Delta {\\bf x}| \\simeq 2\\pi/|\\Delta {\\bf k}|$. On the other hand, the spatial distance between two cells in the ($k-x$) space is at least $|\\Delta {\\bf x}| \\simeq 2\\pi/|\\Delta {\\bf k}|$. This yields the factor $\\delta^K_{\\bf x,x'}$ of eq.(2). Therefore, the phase space correlation function eq.(2) for Gaussian field generally is local. Non-linear evolution via gravitational clustering will lead to a non-Gaussian field that will deviate from the locality of eq.(2), even when the initial field is Gaussian. However, turbulence studies have found that if 1.) the initial perturbations are spatially localized, and 2.) the random fields evolve via a self-similar hierarchical cascade process, the phase space correlation function of the evolved field will still be spatially localized (Greiner, Lipa, \\& Carruthers, 1995, Greiner et al. 1996). The perturbations of modes $\\delta({\\bf k, x})$ and $\\delta({\\bf k', x'})$ with ${\\bf x}\\neq {\\bf x'}$ stay statistically uncorrelated or only very weakly correlated during the self-similar hierarchical cascade evolution. That is, the factor $\\delta^K_{\\bf x,x'}$ of eq.(2) is memorized during the dynamical evolution. Phenomenological models that mimic the hierarchical clustering of the cosmic mass field have similar mathematical structures as the hierarchical cascade models of turbulence. For instance, the fractal hierarchy clustering model of Soneira \\& Peebles (1977) is the same as the $\\beta$ model of turbulence (e.g. Frisch, 1995). The block model of Cole \\& Kaiser (1988) is a special case of the multifractal cascade model (Meneveau \\& Sreenivasan 1987, Pando et al 1998). Hence, these models should also memorize the spatial locality. Recently, the spatial locality has been studied with more realistic dynamical models of gravitational clustering. First, if the weakly nonlinear mass field is given by the Zeldovich approximation, the field is found to be spatially localized if the initial perturbations are Gaussian (Pando, Feng \\& Fang 2001). More recently, this result has been extended to fully nonlinear regime (Feng \\& Fang 2004). Using the halo model of the large scale structure (e.g. Cooray \\& Sheth 2002, and references therein), it has been shown that the evolved mass field is approximately spatially localized if the initial perturbations are Gaussian. Although gravitational coupling is long-term, the spatial locality in the phase space is not disturbed by the non-linear evolution of cosmic mass field. This property essentially is due to the self-similarity of the hierarchical clustering. This result has been tested with high resolution N-body simulation samples (Feng \\& Fang 2004). In this paper, we investigate the spatial locality of phase space correlations with real sample -- the galaxies selected from the 2MASS Extended Source Catalog (XSC, Jarrett et al. 2000). Our motivation is two-fold. First, the spatial locality provides a test of the Gaussianity of the initial density perturbations on small scales. Although many tests on the Gaussianity of the initial perturbations have been done with the temperature fluctuations of the Cosmic Microwave Background Radiation (e.g. Komatsu, et al. 2003, Pando, Valls-Gabaud \\& Fang, 1998), the comoving scales of these tests are not less than a few Mpc. The spatial locality of the 2MASS samples can test the Gaussianity to scale as small as about 0.04 h$^{-1}$ Mpc. Second, the distribution of galaxies is biased from the mass field of dark matter. Some bias models assume that the relation between the distribution of galaxies and underlying mass field is stochastic and nonlocal (e.g. Dekel \\& Lahav, 1999). This mechanism will lead to nonlocality of the galaxy correlation function in phase space, even when the underlying dark matter mass field is spatially localized. Therefore, the spatial locality should be effective in testing the stochasticity and nonlocality of bias models. The outline of this paper is as follows. \\S 2 presents the statistics and dynamics of the spatial locality in phase space. \\S 3 describes the basic properties of the 2MASS galaxy samples with a space-scale decomposition. The results of the spatial locality of the 2MASS galaxy distribution are presented in \\S 4. Finally, the conclusions and discussions are be given in \\S 5. ", "conclusions": "The 2-D distribution of 2MASS galaxies has been studied using the DWT, in which the SFC $\\epsilon_{\\bf j,l}$ is the count-in-cell in physical space, while the WFC $\\tilde{\\epsilon}_{\\bf j,l}$ is a count-in-cell in phase space. We find that the statistical properties of the SFC and WFC variables are very different. The former is non-Gaussian and nonlocally correlated, while the later is non-Gaussian, but its correlation basically is spatially localized. That is, the sample has the following statistical behavior in the phase space \\begin{itemize} \\item the one-point distribution of $\\tilde{\\epsilon}_{\\bf j,l}$ is highly non-Gaussian on angular scales less than 1$^{\\circ}$, corresponding to $\\simeq 3.5$ h$^{-1}$ Mpc at the median redshift of the sample. \\item The 2nd and 4th order mode-mode correlations with $\\tilde{\\epsilon}_{\\bf j,l}$ are always spatially localized. For highest $j$, the locality holds on angular scales as small as 0.06$^{\\circ}$, or $\\simeq$ 0.2 h$^{-1}$ Mpc at the median redshift of the sample. \\item the local scale-scale correlations are significant on scales less than 0.4$^{\\circ}$, corresponding to $\\simeq 1.5$ h$^{-1}$ Mpc at the median redshift of the sample. The scale-scale correlation is also approximately spatially localized. \\end{itemize} A direct physical meaning of these results is that the cosmic gravitational instability causes only strong interaction between the modes in different wavebands and in the same spatial area, but is weak for modes in different spatial area. Because the nonlinear evolution of cosmic gravitational clustering has the memory of its initial spatial correlation in the phase space, the observed spatial locality of the 2MASS galaxies provides solid evidence for models assuming that the initial perturbations are spatially uncorrelated among phase space modes. This result is consistent with the assumption that the initial perturbations are Gaussian. Although the Gaussianity of the initial perturbations on large scales has been extensively tested with the temperature fluctuations of the cosmic microwave background radiation, the test given by the spatial locality is effective to comoving scales as small as $\\simeq$ 0.2 h$^{-1}$ Mpc. The spatial locality has been studied with samples of Ly-alpha forests (Pando, Feng \\& Fang, 2001). Since Ly-alpha forests refer to weakly non-linear clustering, their spatial locality can be used to rule out the models of initially non-Gaussian fields which are non-spatially localized, but cannot do so for models of initial non-Gaussian which are spatially localized. In this paper, the spatial locality is found for sample referring to fully nonlinear regime. The memory of the spatial locality is based on the halo model, for which an initially Gaussian field is necessary. In this sense, the spatial locality of the 2MASS gives stronger support to the initially Gaussian assumption than that that of Ly-alpha forests. The one-point distribution of 2MASS galaxies (Fig. 3) is very different from that given by N-body simulation sample. The later in nonlinear regime generally is lognormal (see Fig. 6 of Feng \\& Fang, 2004), while the former is more complicated (Fig. 3). This difference indicates that galaxy distribution is biased with respect to the underlying dark matter field. Nevertheless, the galaxy distribution is also highly spatially localized. This result indicates that relationship between $\\rho_g({\\bf x})$ and $\\rho({\\bf x})$ should be localized in phase space. That is, the stochasticity and nonlocality of the relation between $\\rho_g({\\bf x})$ and $\\rho({\\bf x})$ are limited in the cell of phase space. For a structure consisting of perturbations on scales ${\\bf k}$ to ${\\bf k}+\\Delta {\\bf k}$, the nonlocal size in the relation between $\\rho_g({\\bf x})$ and $\\rho({\\bf x})$ should not be larger than $|\\Delta {\\bf x}|=2\\pi/|\\Delta {\\bf k}|$. Otherwise, the galaxy bias will violate the spatial locality of the galaxy distribution even when the underlying field is spatial localized. Hence, from the 2MASS galaxies we can conclude that the stochasticity and nonlocality of the bias relation between $\\rho_g({\\bf x})$ and $\\rho({\\bf x})$ probably are no more than that given by the uncertainty relation $|\\Delta {\\bf x}||\\Delta {\\bf k}|=2\\pi$." }, "0403/astro-ph0403691_arXiv.txt": { "abstract": "We have fitted the {\\it Chandra\\/} Low Energy Transmission Grating spectrum of SS~Cygni in outburst with a single temperature blackbody suffering the photoelectric opacity of a neutral column density and the scattering opacity of an outflowing wind. We find that this simple model is capable of reproducing the essential features of the observed spectrum with the blackbody temperature $T_{\\rm bl}\\approx 250\\pm 50$ kK, hydrogen column density $N_{\\rm H}\\approx 5.0^{+2.9}_{-1.5}\\times 10^{19}~\\rm cm^{-2}$, fractional emitting area $f\\approx 5.6^{+60}_{-4.5}\\times 10^{-3}$, boundary layer luminosity $L_{\\rm bl}\\approx 5^{+18}_{-3}\\times 10^{33}~\\rm erg~s^{-1}$, wind velocity $v\\approx 2500~\\rm km~s^{-1}$, wind mass-loss rate $\\Mdot_{\\rm w}\\approx 1.1\\times 10^{16}~\\rm g~s^{-1}$, and arbitrary values of the wind ionization fractions of 20 ions of O, Ne, Mg, Si, S, and Fe. Given that in outburst the accretion disk luminosity $L_{\\rm disk}\\approx 1\\times 10^{35}~\\rm erg~s^{-1}$, $L_{\\rm bl}/L_{\\rm disk}\\approx 0.05^{+0.18}_{-0.03}$, which can be explained if the white dwarf (or an equatorial belt thereon) is rotating with an angular velocity $\\Omega_{\\rm wd}\\approx 0.7^{+0.1}_{-0.2}$ Hz, hence $V_{\\rm rot}\\sin i\\sim 2300~\\rm km~s^{-1}$. ", "introduction": "According to simple theory, the boundary layer between the accretion disk and the surface of the white dwarf is the dominant source of high energy radiation in nonmagnetic cataclysmic variables (CVs). Unless the white dwarf is rotating rapidly, the boundary layer luminosity $L_{\\rm bl}\\approx G\\Mwd\\Mdot /2\\Rwd $, where $\\Mwd $ and $\\Rwd $ are respectively the mass and radius of the white dwarf and $\\Mdot $ is the mass-accretion rate. When $\\Mdot $ is low, as in dwarf nova in quiescence, the boundary layer is optically thin to its own radiation and its temperature is of order the virial temperature $T_{\\rm vir} = G\\Mwd m_{\\rm H}/3k\\Rwd\\sim 10$ keV. When $\\Mdot $ is high, as in nova-like variables and dwarf novae in outburst, the boundary layer is optically thick to its own radiation and its temperature is of order the blackbody temperature $T_{\\rm bb} = (G\\Mwd\\Mdot /8\\pi\\sigma f\\Rwd^3)^{1/4}\\sim 100$ kK. The compact nature of the boundary layer is established by the narrow eclipses of the hard X-ray flux of high inclination dwarf novae in quiescence \\citep{woo95a, muk97, tes97, pra99a, whe03b} and the short periods [$P\\sim 10~{\\rm s}\\sim 2\\pi(\\Rwd ^3/G\\Mwd )^{1/2}$] of the oscillations of the soft X-ray and extreme ultraviolet (EUV) flux of nova-like variables and dwarf novae in outburst \\citep{war04, mau04a}. However, an {\\it extended\\/} source of soft X-ray and EUV flux is required to explain the lack of eclipses in high inclination nova-like variables and dwarf novae in outburst \\citep{nay88, woo95b, pra99b, mau00, pra04}. Using the {\\it Extreme Ultraviolet Explorer\\/} ({\\it EUVE\\/}) light curve and spectrum of OY Car in superoutburst, \\citet{mau00} argued that the source of the extended EUV flux of high-$\\Mdot $ nonmagnetic CVs is the accretion disk wind. In this picture, in high inclination systems the boundary layer is hidden by the accretion disk for all binary phases, but its flux is scattered into the line of sight by resonance transitions of ions in the wind. Because the scattering optical depths of these transitions are much higher than the Thompson optical depth, the EUV spectra of such systems should be dominated by broad lines. Consistent with this picture, in outburst the EUV spectra of the high inclination dwarf novae OY~Car ($i\\approx 83^\\circ $; \\citealt{mau00}) and WZ~Sge ($i\\approx 75^\\circ $; Wheatley \\& Mauche 2004, in preparation) are dominated by broad lines, while that of the low inclination dwarf novae VW~Hyi ($i\\sim 60^\\circ $; \\citealt{mau96}) and SS~Cyg ($i\\sim 40^\\circ $; \\citealt{mau95}) are dominated by the continuum. U~Gem ($i\\approx 70^\\circ $; \\citealt{lon96}) is a transitional case, as its EUV flux is partially eclipsed and its EUV spectrum contains of broad emission lines superposed on a bright continuum. The general case of the radiation transfer of the boundary layer flux through the accretion disk wind is complicated, but schematically we expect that in high inclination systems the EUV spectrum will be of the form $F_\\lambda\\, e^{-\\tau_\\lambda ^\\prime}\\, (1-e^{-\\tau_\\lambda})$, while in low inclination systems it will be of the form $F_\\lambda\\, e^{-\\tau_\\lambda ^\\prime} e^{-\\tau_\\lambda}$, where $F_\\lambda$ is the intrinsic boundary layer spectrum, $\\tau_\\lambda ^\\prime$ is the photoelectric optical depth of the wind and the interstellar medium, and $\\tau_\\lambda$ is the scattering optical depth of the wind. \\cite{mau00} showed that the second version of this model describes the main features of the {\\it EUVE\\/} spectrum of OY~Car in superoutburst, and we show here that the first version of this model describes the main features of the {\\it Chandra\\/} Low Energy Transmission Grating (LETG) spectrum of SS~Cyg in outburst. A preliminary discussion of this spectrum is supplied by \\citet{mau04a}, while the quasi-coherent oscillations of the EUV flux is discussed by \\citet{mau02}. ", "conclusions": "We have fitted the {\\it Chandra\\/} LETG spectrum of SS~Cyg in outburst with a single temperature blackbody suffering the photoelectric opacity of a neutral column density and the scattering opacity of an outflowing wind. The best-fit model, shown in Figure~3, has parameters $T_{\\rm bl}=250$ kK, $N_{\\rm H} = 5.0\\times 10^{19}~\\rm cm^{-2}$, $f=5.6\\times 10^{-3}$, $\\Mdot_{\\rm w}=1.1 \\times 10^{16}~\\rm g~s^{-1}$, and the values of $X$ listed in Table~1 ($X=1$ otherwise). Assuming that the wind is optically thin and photoionized by a 250 kK blackbody, the range of ions present in the model requires that the ionization parameter $-1\\lax\\log\\xi\\lax 5$. Given these results, in outburst the boundary layer luminosity $L_{\\rm bl} = 4.7\\times 10^{33}~\\rm erg~s^{-1}$. A lower limit on the value of this quantity is imposed by the value of the interstellar H column density, which \\citet{mau88} estimated to be $3.5\\times 10^{19}~\\rm cm^{-2}$ based on the curve-of-growth of interstellar absorption lines in high resolution {\\it International Ultraviolet Explorer\\/} spectra of SS~Cyg in outburst. For $-1\\lax\\log\\xi\\lax 5$, H and He are fully ionized, so the wind should have no photoelectric opacity in the EUV bandpass, and $N_{\\rm H}$ should be equal to the interstellar H column density. With $N_{\\rm H}=3.5\\times 10^{19}~\\rm cm^{-2}$ and $T_{\\rm bl}=300$ kK, Figure~2 shows that $f=1.1\\times 10^{-3}$, hence $L_{\\rm bl}=1.7\\times 10^{33}~\\rm erg~s^{-1}$. At the other end of the $\\chi^2$ ellipse, with $T_{\\rm bl}=200$ kK and $N_{\\rm H}=7.9\\times 10^{19}~\\rm cm^{-2}$, $f=6.6\\times 10^{-2}$, hence $L_{\\rm bl}= 2.3\\times 10^{34}~\\rm erg~s^{-1}$. We conclude that a reasonable estimate of the boundary layer luminosity $L_{\\rm bl}\\approx 5^{+18}_{-3}\\times 10^{33}\\, (d/160~{\\rm pc})^2~\\rm erg~s^{-1}$. How does this compare with the accretion disk luminosity? \\citet{pol84} used optical through far ultraviolet spectra of SS~Cyg in outburst to determine that $L_{\\rm disk}\\approx 1\\times 10^{35}\\, (d/160~{\\rm pc})^2~\\rm erg~s^{-1}$, so $L_{\\rm bl}/L_{\\rm disk}\\approx 0.05^{+0.18}_{-0.03}$. Theoretically, we expect that this ratio is equal to one unless the white dwarf is rotating rapidly, in which case $L_{\\rm bl}/L_{\\rm disk}=(1-\\omega)^2$, where $\\omega =\\Omega_{\\rm wd}/\\Omega_{\\rm K}(\\Rwd )$, $\\Omega_{\\rm wd}=2\\pi/P_{\\rm spin}$ is the angular velocity of the white dwarf, and $\\Omega_{\\rm K}(\\Rwd ) = (G\\Mwd/\\Rwd^3)^{1/2}\\approx 0.9$ Hz is the Keplerian angular velocity of material just above its surface \\citep{pop95}. The white dwarf can be spun up to the required rate by accreting an amount of mass $\\Delta M=\\Mwd - M_{\\rm wd,i}$, where $M_{\\rm wd,i}$ is the initial white dwarf mass, $\\Mwd = M_{\\rm wd,i}\\, (1-4k^2\\omega/3)^{-3/4}$ is the final white dwarf mass, and $k$ is the radius of gyration \\citep{lan00}. With $k\\approx 0.4$ and $\\omega = 1- (L_{\\rm bl}/L_{\\rm disk})^{1/2}\\approx 0.78^{+0.08}_{-0.26}$, $\\Delta M\\approx 0.13\\, \\Mwd\\sim 0.1~\\Msun $. If rapid rotation is the cause of SS~Cyg's low boundary layer luminosity, the white dwarf angular velocity $\\Omega_{\\rm wd} =\\omega\\Omega_{\\rm K}(\\Rwd )\\approx 0.7^{+0.1}_{-0.2}$ Hz, hence the spin period $P_{\\rm spin}\\approx 9^{+4}_{-1}$ s (comparable to the period of the quasi-coherent oscillations), the rotation velocity $V_{\\rm rot} = \\Omega_{\\rm wd}\\Rwd\\approx 3800^{+400}_{-1300}~\\rm km~s^{-1}$, and $V_{\\rm rot}\\sin i\\sim 2300~\\rm km~s^{-1}$. This last quantity is rather large, given that $V_{\\rm rot}\\sin i\\approx 1200~\\rm km~s^{-1}$ for WZ~Sge and $V_{\\rm rot}\\sin i\\lax 400~\\rm km~s^{-1}$ for six other nonmagnetic CVs \\citep{sio99}. Finally, we note that in the single scattering limit, conservation of momentum limits the mass-loss rate of a radiatively driven wind to $\\Mdot_{\\rm max} =L/V_\\infty c$, where $L$ is the system luminosity and $V_\\infty $ is the wind terminal velocity. For SS~Cyg in outburst, $L=L_{\\rm disk}+L_{\\rm bl}\\approx 1\\times 10^{35}~\\rm erg~s^{-1}$ and $V_\\infty\\approx 3000~\\rm km~s^{-1}$, so $\\Mdot_{\\rm max}\\approx 1\\times 10^{16}~\\rm g~s^{-1}$. We derived a wind mass-loss rate $\\Mdot_{\\rm w}\\approx 1\\times 10^{16}~\\rm g~s^{-1}$ for SS~Cyg in outburst, so $\\Mdot_{\\rm w}/\\Mdot_{\\rm max}\\approx 1$. In $\\tau $~Sco (B0 V), P~Cyg (B1 Ia$^+$), $\\zeta $~Pup (O4 f), and $\\epsilon $~Ori (B0 Ia) $\\Mdot_{\\rm w}/\\Mdot_{\\rm max}=0.02$, 0.22, 0.33, and 0.65, but in WR1 (WN5) $\\Mdot_{\\rm w}/\\Mdot_{\\rm max}=60$ \\citep{lam99}, so we cannot conclude from this result that a mechanism other than radiation pressure is needed to drive the wind of SS~Cyg." }, "0403/astro-ph0403372_arXiv.txt": { "abstract": "In the last few years new evidence has been presented for the presence of ongoing massive star formation in the outer HI disks of galaxies. These discoveries strongly suggest that precursor molecular gas must also be present in some physical state which is escaping detection by the usual means (CO(1-0), IR, etc.). We present a model for such a gas in a framework which views the HI as the result of an ongoing ``photodissociation $\\Leftrightarrow$ dust grain reformation'' equilibrium in a cold, clumpy molecular medium with a small area filling factor. ", "introduction": "Recently Komiyama et al.\\ (2003) and De Blok \\& Walter (2003) have reported the detection of young stars in the outer HI disk of the nearest dwarf irregular galaxy NGC 6822 ($D \\approx 490$ kpc) at galactocentric radii well beyond $R_{25}$. These authors discuss the results in terms of the commonly-accepted picture, where young massive stars form out of the observed HI in response to some ``triggering'' mechanism such as the passage of a companion galaxy (in this case a massive intergalactic HI cloud). However, an ``interaction'' scenario is problematic for NGC 6822, since as De Blok \\& Walter point out there is little evidence for tidal ``impact trauma'' on the distribution of old stars in the galaxy. ", "conclusions": "\\begin{itemize} \\item \\textit{\\textbf{The hypothesis:} An $H_2$ photodissociation $\\Leftrightarrow$ reformation equilibrium scenario provides a simple explanation for the coexistence of young stars, dust, and HI in the outer parts of disk galaxies.} \\item \\textit{\\textbf{The hypothesis is falsifiable:} If we find dynamically quiescent HI regions in the outer parts of galaxies without dissociating stars ($M(V) \\leq -1$, B5 or brighter, $M \\geq 6 M_{\\sun}$) then this idea is wrong.} \\item \\textit{\\textbf{A prediction:} Even low surface brightness galaxies such as NGC 2915 will have dissociating B-stars mixed in with their HI.} \\end{itemize}" }, "0403/astro-ph0403144_arXiv.txt": { "abstract": "A modified non-linear time series analysis technique, which computes the correlation dimension $D_2$, is used to analyze the X-ray light curves of the black hole system GRS 1915+105 in all twelve temporal classes. For four of these temporal classes $D_2 $ saturates to $\\approx 4-5$ which indicates that the underlying dynamical mechanism is a low dimensional chaotic system. Of the other eight classes, three show stochastic behavior while five show deviation from randomness. The light curves for four classes which depict chaotic behavior have the smallest ratio of the expected Poisson noise to the variability ($ < 0.05$) while those for the three classes which depict stochastic behavior is the highest ($ > 0.2$). This suggests that the temporal behavior of the black hole system is governed by a low dimensional chaotic system, whose nature is detectable only when the Poisson fluctuations are much smaller than the variability. ", "introduction": "\\label{sec: I} Black hole X-ray binaries are variable on a wide range of timescales ranging from months to milli-seconds. A detailed analysis of their temporal variability is crucial to the understanding of the geometry and structure of these high energy sources. Such studies may eventually be used to test the relativistic nature of these sources and to understand the physics of the accretion process. The variability in different energy bands is generally quantified by computing the power spectrum which is the amplitude squared of the Fourier transform. The power spectra give information about the characteristic frequencies of the system which show up as either breaks or as near Gaussian peaks, i.e Quasi-Periodic Oscillation (QPO) in the spectra (e.g. \\citet{Bel01,Tom01,Rod02}). The shape of the power spectra, combined with the observed frequency dependent time lags between different energy bands, have put constraints on the radiative mechanisms and geometry of emitting regions (e.g. \\citet{Now99,Mis00,Cui99,Pou99,Cha00,Nob01}) These results are based on the response of the system to temporal variations whose origin is not clear. Important insight into the origin can be obtained by the detection and quantification of the possible non-linear behavior of the fluctuations. For example, the presence of stochastic fluctuations would favor X-ray variations driven by variations of some external parameters (like the mass accretion rate), or the possibility that active flares occur randomly. On the other hand, if the fluctuations can be described as a deterministic chaotic system, then inner disk instability or coherent flaring activity models will be the likely origin. A quantitative description of the temporal behavior can also be compared with time dependent numerical simulations of the accretion process and will help examine the physical relevance of these simulations. The non-Gaussian and non-zero skewness values of the temporal variation of the black hole system Cygnus X-1 suggested that the variations are non-linear in nature \\citep{Thi01,Tim00,Mac02}. More rigorous tests were applied to the AGN ArK 564 \\citep{Gli02} which also suggested non-linear behavior. Nonlinear time series analysis(NLTS) seems to be the most convenient tool to check if the origin of the variability is chaotic, stochastic or a mixture of the two and has been adopted in several disciplines to study complex systems (e.g. human brain, weather) and predict their immediate future \\citep{Sch99}. This technique has also been used earlier to analyze X-ray data of astrophysical sources. Based on a NLTS analysis of EXOSAT data, \\cite{Vog87} claimed that the X-ray Pulsar Her X-1 was a low dimensional chaotic system. However, \\cite{Nor89} pointed out problems with that analysis since the source has a strong periodicity and the data analyzed had low signal to noise ratio. \\cite{Leh93} used the NLTS technique to analyze EXOSAT light curves of several AGN, and found that only one, NGC 4051, showed signs of low dimensional chaos. A similar analysis on the noise filtered{\\it Tenma} satellite data of Cyg X-1, suggested that the source may be a low dimensional chaotic system with large intrinsic noise \\citep{Unn90}. These analysis were hampered by small number of data points ($\\simless 1000$) in the light curve and/or noise. Hence, the reported detection of low dimension chaos was only possible by rather subjective comparison of the results of the data analysis with those from simulated data of chaotic systems with noise. The Galactic micro quasar GRS 1915+105 is a highly variable black hole system. It shows a wide range of variability \\citep{Che97,Pau97,Bel97a} which required \\citet{Bel00} to classify the behavior in no less than twelve temporal classes. In this work, our motivation is to determine the temporal property of this source by using a modified nonlinear time series analysis for each of these twelve classes. The different kinds of variability and its brightness ( the average RXTE PCA count rate ranges from $5000-32000$ counts/s) makes this source an ideal one to detect chaotic behavior. In the next section we describe the technique used to determine the Correlation dimension. The results of the analysis are presented in \\S 3, while in \\S 4 the work is summarized and discussed. ", "conclusions": "The saturation of the correlation dimension $D_2 \\approx 4-5$ for four of the temporal classes clearly indicates that the underlying dynamic mechanism that governs the variability of the black hole system is a low dimensional chaotic one. As indicated by simulations of the Lorenz system with noise, the effect of Poisson noise in the data is to increase the $D_2$ values. Hence the real dimension of the system is probably smaller than $D_2 \\approx 4-5$ that is obtained here. In fact it is possible that the the temporal behavior of the black hole system is always governed by a low dimensional chaotic system, but is undetectable when Poisson noise affects the analysis. Alternatively, there may be a stochastic component to the variability which dominates for certain temporal classes. The two scenarios may be distinguished and better quantitative estimates of the correlation dimension may be obtained by either appropriate noise filtering of the data and/or appropriate averaging of the different light curves. Much longer ($\\approx 30000$ sec long) continuous data streams sampled at $1$ second resolution, would decrease Poisson noise and hence provide better quantitative measure of $D_2$. However, such long data streams are presently not available and merging non-continuous light curves, will require sophisticated gap filling techniques which might give rise to spurious results. The variability of GRS 1915+105 can be interpreted as being transitions between three spectral states \\citep{Bel00}, one of which (the so called soft state) is a long term canonical state observed in other black hole systems like Cygnus X-1 which do not show such high amplitude variability. It is attractive to identify these spectral states as fixed points which for GRS 1915+105 become unstable giving rise to the observed chaotic behavior which may also account for the ring like movement of the system in color-color space \\citep{Vil98}. The above hypothesis may be verified by future characterization of the chaos in GRS 1915+105. Note that GRS15+105 spends most of it's time in the $\\chi$ class whose variability is similar to that observed in other black hole systems like Cygnus X-1. However, as shown in this work, Poisson noise effects the analysis for the $\\chi$ class and the $D_2 (M)$ values reflect stochastic behavior. This may be the reason why earlier different non-linear analysis of Cygnus X-1 data, while showing non-linearity \\citep{Tim00,Thi01} did not conclusively reveal chaotic behavior. The identification of the temporal behavior of the black hole system as a chaotic one, has opened a new window toward the understanding of the origin and nature of their variability. The present analysis can be extended to characterize the chaotic behavior. Using the minimum required phase space dimension, the data can be projected into different $2$ dimensional planes, which will reveal the structure of the attractor and help to identify any possible centers of instability in the system. Further, dynamical invariants like the full Lyapunov spectrum , multi-fractal dimensions etc. can be also be computed. Recently, \\citet{Win03} have studied and quantified the chaotic flow in magneto-hydrodynamic simulations of the mass accretion processes that is believed to be happening in black hole systems. The measured chaos parameters like the largest Lyapunov exponent for such simulations can be compared with that obtained from the light curve of black hole systems to validate such simulations and enhance our understanding of these systems. Note that such analysis can practically be applied only after the identification of the minimum phase space dimension which in turn usually requires the computation of $D_2 (M)$." }, "0403/astro-ph0403199_arXiv.txt": { "abstract": "Mean pulse profiles and polarization properties are presented for nine southern pulsars. The observations were made using the Parkes radio telescope at frequencies near 1330 MHz; three of the nine pulsars were also observed at 660 MHz. A very high degree of circular polarization was observed in PSR J1603$-$7202. Complex position angle variations which are not well described by the rotating-vector model were observed in PSRs J2124$-$3358 and J2145$-$0750, both of which have very wide profiles. Rotation measures were obtained for all nine pulsars, with two implying relatively strong interstellar magnetic fields. ", "introduction": "Observations of the polarization of mean pulse profiles have given much information about the pulsar radio emission mechanism. Pulsars often have very high linear polarization and the sweep of position angle across the pulse is generally well described by the rotating-vector model \\citep{rc69a,kom70}. This and the bilateral symmetry of many pulse profiles led to a model in which the emission is beamed into a hollow cone centered on the magnetic axis. With observations of an increasing number of pulsars it soon became clear that the mean pulse profiles of many pulsars were more complex. The frequent occurence of a central emission peak \\citep{bac76} led to the idea of a `core' component in contrast to the outer `conal' components. Further observations showed that many pulsars had multiple conal components leading to ideas of multiple cones \\citep{ran93,kwj+94,md99} or patchy beams \\citep{lm88,hm01}. Core emission tends to have a steeper spectrum than conal emission and a higher degree of circular polarization, often with a reversal of sense near the profile center \\citep{ran83,lm88}, although reversing circular polarization is sometimes observed in conal emission \\citep{hmxq98}. The apparent width of both core and conal beams is a function of pulse period ($P$), with an approximate $P^{-1/2}$ dependence \\citep{ran90,big90b}. With the discovery of millisecond pulsars (MSPs), the parameter ranges available for study of pulse emission processes was greatly extended. MSPs have much weaker surface magnetic fields than `normal' pulsars, with $B_0$ typically a few by $10^8$~G, compared to $10^{12}$~G for normal pulsars. They also have much smaller magnetospheres, conventionally defined to be limited by the light cylinder at radius $R_{LC} = cP/2\\pi$, where $P$ is the pulsar period and $c$ is the velocity of light. For the fastest MSP, PSR B1937+21, $R_{LC}$ is only 74 km, or about ten times the neutron star radius. One would therefore expect the radio emission from MSPs to be rather different to that from normal pulsars. MSP pulse profiles are generally wider (in pulse phase) than those of normal pulsars, which is not surprising in view of the wider angle subtended by the `open' field lines which penetrate the light cylinder and define the extent of the polar cap. Also, MSP profiles have different frequency evolution characteristics compared to most normal pulsars. In particular, the component separations are largely independent of frequency \\citep{kll+99}. It is not clear that the distinctions and relationships between `core' and `conal' emission seen in longer-period pulsars apply to MSPs. For example, the strong component in PSR J0437$-$4715 is clearly central to the emission beam but has a flatter spectrum than the outer components \\citep{nms+97}. Despite these differences, there are many similarities in the radio emission properties of MSPs and normal pulsars. If the greater width of MSP profiles is ignored, the shapes of MSP mean pulse profiles are very similar to those of normal pulsars. However, among the MSPs, there is a greater proportion of `interpulse' profiles, with two main regions of pulse emission separated by approximately $180\\degr$ of pulse phase or longitude. MSP profiles often appear to have more components than normal pulsars, but Kramer et al. (1998)\\nocite{kxl+98} have argued that this is simply because MSP components are more obvious because of their wider spacing. On average, the radio emission of MSPs has a spectral index which is only marginally steeper than that of normal pulsars and may in fact be the same \\citep{tbms98,kxl+98}. MSPs are typically somewhat less luminous than normal pulsars but there is a large overlap in the luminosity distributions. Finally, the polarization properties of MSPs and normal pulsars are remarkably similar, with both often having high fractional linear polarization and generally smaller levels of circular polarization. Orthogonal jumps in position angle are seen in both classes of pulsar (e.g. Stinebring et al. 1984\\nocite{scr+84}; Thorsett \\& Stinebring 1990\\nocite{ts90}). Variations of position angle (PA) through MSP profiles are often more complex than is typical for normal pulsars. Despite this, in many cases the PA can be approximately fitted by the simple rotating vector model (RVM) which applies to most normal pulsars, at least once orthogonal jumps are taken into account (e.g. Arzoumanian et al. 1996)\\nocite{aptw96}. These similarities leave little doubt that the basic radio emission mechanism is the same in normal pulsars and MSPs. Extensive studies of the polarization properties of MSPs have previously been made by Thorsett \\& Stinebring (1990)\\nocite{ts90}: PSRs B1937+21, B1953+29 and B1957+20 at several frequencies between 430 and 2830 MHz; Navarro et al. (1997)\\nocite{nms+97}: PSR J0437$-$4715 at 438, 660 and 1512 MHz; Arzoumanian et al. (1996)\\nocite{aptw96}: PSR B1534+12 at 430 MHz; Xilouris et al. (1998)\\nocite{xkj+98}: 24 pulsars at 1410 MHz; Stairs, Thorsett \\& Camilo (1999)\\nocite{stc99}: 19 pulsars at 410, 610 and/or 1414 MHz; and Lommen et al. (2000)\\nocite{lzb+00}: PSR J0030+0451 at 430 MHz. All observations except those of PSR J0437$-$4715 by Navarro et al. were made using northern hemisphere telescopes, so there is little information on the polarization of the southern MSPs. Also, there are significant discrepancies in the results of different groups for some pulsars. For example, results from Xilouris et al. (1998) and Stairs et al. (1999) for PSRs J1022+1001, J1713+0747 and J2145-0750 at frequencies near 1400 MHz differ substantially.\\footnote{Stairs et al. (1999) use a convention for sense of position angle which is opposite to the astronomical convention used by other authors.} It is important to understand whether these differences are due to time variations such as mode changing \\citep{kxc+99} or problems with data processing and/or calibration. Furthermore, in some previous studies (e.g. Stairs et al. 1999) the position angle scale is not absolute. Absolute position angles are important for Faraday rotation studies and in comparisons with other properties. Important examples are the comparison of the pulsar rotation axis direction implied by radio polarization measurements with that deduced from X-ray observations and with the direction of the pulsar proper motion. \\citep{hgh01,rd01}. With their high linear polarization, it is relatively easy to determine rotation measures (RMs) for most pulsars. Combined with the dispersion measure (DM), pulsars give a direct measure of the mean line-of-sight magnetic field between the Sun and the pulsar, weighted by the local electron density. Since the DM also provides an estimate of the pulsar distance, pulsars are a valuable probe of the Galactic magnetic field (e.g. Han et al. 2002)\\nocite{hmlq02}. Most MSPs are relatively close to the Sun, so they provide information on conditions in the local region of the Galaxy. We have used the Parkes 64-m telescope of the Australia Telescope National Facility and the Caltech pulsar correlator to make observations of nine southern MSPs at 20~cm and, for three of these, at 50~cm. Five of these pulsars have no previously published polarization observations and none have previously published rotation measures. Details of the observing system and observational parameters are given in \\S~2. The polarization and RM results are presented in \\S~3 and \\S~4 respectively, and in \\S~5 we discuss some implications of our polarization measurements. ", "conclusions": "In this paper we have presented pulse profiles and polarization parameters for nine southern MSPs, five of which had no previously published polarization data. New RM values are given for all nine pulsars. Notable results include the very high fractional circular polarization ($\\sim 65$\\% at 1327 MHz) in the trailing component of PSR J1603$-$7202, and the extremely wide and complex pulse profile and polarization properties of PSR J2124$-$3358. \\citet{kxl+98} have argued that, on average, MSPs have profiles which are of similar complexity (quantified by the number of recognizable components) to those of normal pulsars. However, there is a tail to the distribution in MSPs exemplified by PSR J0473$-$4715 \\citep{nms+97}, which has at least 10 components, and PSR J2124$-$3358 (Figure~\\ref{fg:2124}; at least 12 components) which is not present in normal pulsars; no normal pulsars have profiles as complex as these, even in data with high signal-to-noise ratio. While many MSP profiles have the appearance of `stretched' normal pulsar profiles, there are significant differences. One of the most important is the lack of frequency evolution in the spacing of pulse components \\citep{kll+99}. In most normal pulsars, the pulse components are more widely separated at lower radio frequencies, an observation often interpreted in terms of `radius-to-frequency mapping' \\citep[e.g.][]{cor78,tho91a}. The results presented here confirm this lack of frequency evolution in MSP profiles. For example, Figure~\\ref{fg:2124_alpha} shows that for PSR J2124$-$3358, although components are mostly overlapping, there is a clear correspondence of component locations at the two frequencies right across the profile. It is important to note however that this property is shared by normal pulsars with so-called `interpulse' emission, including such well-known examples as the Crab pulsar \\citep{mh99} and PSR B0950+08 \\citep{hc81}, suggesting a close link between the emission processes in MSPs and `normal' interpulse pulsars. Most MSP profiles do not fit easily into the `core-cone' classification scheme \\citep{ran83,lm88} or its extensions to multiple cones \\citep[e.g.][]{md99}. Even when the profile has a shape similar to a classical `triple', for example PSR J2145$-$0750 (Figure~\\ref{fg:2145}), the spectral properties of the central and outer components do not follow the patterns established for normal pulsars. For PSR J2124$-$3358 there is no identifiable `core' component and there is no pattern suggesting multiple nested conal emission. While there do appear to be discrete quasi-gaussian components within the observed profile, a very large number of them would be required to model it and their amplitude and phase (longitude) shows no clear pattern. Rather, the profile seems more consistent with a random distribution of emission patches as advocated by \\citet{lm88}. The spectral index variations shown in Figure~\\ref{fg:2124_alpha} are consistent with different pulse components originating in spatially distinct emission regions. The wide observed profiles of MSPs give a good opportunity to fit the RVM and hence to determine both the magnetic inclination $\\alpha$ and the line-of-sight inclination $\\zeta$. Most normal pulsars have narrow profiles and only the impact parameter $\\beta = \\zeta - \\alpha$ can be measured \\citep[cf.][]{lm88}. In some MSPs, the observed PA variations follow the RVM quite well, especially at higher frequencies \\citep[e.g.][]{stc99,lzb+00}. However, in others, there are clear and sometimes large deviations \\citep[e.g.][]{aptw96,nms+97}. Two pulsars in the present sample, PSRs J2124$-$3358 and 2145$-$0750, have sufficiently wide profiles to justify an attempt to fit the RVM. The results are shown in Figure~\\ref{fg:pafits}. \\begin{figure*} \\psfig{file=f11.eps,width=160mm,angle=270} \\caption{Rotating-vector model fits to observed position-angle variations for two pulsars.} \\label{fg:pafits} \\end{figure*} While the overall PA variation for PSR J2124$-$3358 is approximately described by the RVM fit, it is clear that there are large and systematic deviations from the model fit. The parameters of this fit are $\\alpha = 48\\degr\\pm 3\\degr$ and $\\zeta = 67\\degr\\pm 5\\degr$, suggesting that emission is seen from both poles with impact parameters $\\beta = 19\\degr$ and $65\\degr$ respectively. However, there is a large covariance between $\\alpha$ and $\\zeta$ with only a small increase in $\\chi^2$ for $\\alpha$ values between $20\\degr$ and $60\\degr$. At these limits, the corresponding $\\zeta$ values are $27\\degr$ and $80\\degr$ respectively. For small $\\alpha$, a one-pole interpretation is favoured. Given the poor quality of the fit, it is not at all clear what the true inclination angles for the two axes are. Certainly, the fact that emission is seen over most if not all of the pulse period suggests a one-pole, almost aligned model. For PSR J2145$-$0750, if we assume there is an orthogonal flip near longitude $120\\degr$ but not at $195\\degr$, we get the fit shown in Figure~\\ref{fg:pafits}. This has $\\alpha = 145\\degr\\pm 12\\degr$ and $\\zeta = 148\\degr\\pm 12\\degr$. Again, there are systematic deviations from the fit and a large covariance between $\\alpha$ and $\\zeta$, but the impact parameter $\\beta \\simeq 3\\degr$ is relatively well determined. This PA fit, the sense reversal of the circular polarization and the spectral behaviour all suggest that, despite the profile morphology, the trailing component is in fact central to the emission beam \\citep[cf.][]{stc99}. If we assume a second orthogonal jump at $195\\degr$ the impact parameter increases to about $15\\degr$. However, this introduces a significant discontinuity in PA at this longitude, suggesting that this is not the correct interpretation. If we accept that the trailing component is in fact central to the emission beam, the interpretation of the weak leading component becomes something of a problem. \\citet{xkj+98} suggest that it is a `precursor', analogous to that in the Crab pulsar. This is supported by its nearly 100\\% linear polarization. The emission bridge between the leading and strongest components suggests that all components are closely related and probably that all the emission orginates from one pole. This implies a very large effective pulse width, $2\\Delta\\phi \\sim 300\\degr$. For small $\\beta$, $\\sin(\\rho/2) \\simeq \\sin(\\Delta\\phi/2)\\sin(\\alpha/2)$, where $\\rho$ is the true beam radius \\citep[cf.][]{lm88}, giving $\\rho \\sim 68\\degr$. This compares with a predicted value of $49\\degr$ from the relation $\\rho = 6\\fdg2\\;P^{-1/2}$ \\citep{big90b}. For other pulsars in the sample, and indeed for most MSPs \\citep{xkj+98,stc99}, there is very little swing in PA across the the observed pulse components. Taken at face value, this implies large impact parameters and/or very small magnetic inclinations. Yet many of these pulsars have relatively narrow pulses, e.g. PSRs J0621+1002, J1518+4904 and J1713+0747 \\citep{stc99}. This implies a beam radius in the longitude direction much less than the apparent impact parameter, that is, a beam effectively elongated in the latitude direction \\citep[cf.][]{nv83}. If these pulse components are part of a wide beam from magnetic field lines associated with one pole, as suggested for young `interpulse' pulsars by \\citet{man96}, intermediate magnetic inclinations and large impact parameters are possible. In this case, the implied emission altitudes relative to the light cylinder are much higher than those implied by observed pulse widths for normal pulsars. At these altitudes, there will be significant deviations of the magnetic field from a pure dipole form due to displacement and physical currents, providing a possible reason for the observed large PA deviations from RVM fits seen in millisecond pulsars. It is also possible that these irregular PA variations result from very different magnetic field structures in MSPs. For example, \\citet{rud91b} suggests that, during the spindown and subsequent recycling process, crustal plate movements result in highly distorted field structures. Accretion-induced field decay \\citep[e.g.][]{rom90} would also result in highly non-dipolar fields. Such field structures could also disguise or destroy the relatively consistent patterns of core and conal emission seen in most longer-period pulsars. In general, our polarization results agree well with those of \\citet{stc99} when account is taken of the reversed PA convention in that paper. However, there are many discrepancies with results in the \\citet{xkj+98} paper. It appears that their sign of Stokes $V$ is systematically reversed. It is also probable that the higher and approximately constant fractional linear polarization with nearly constant PA seen in a number of their figures results from errors in calibration rather than time variations in the polarization properties \\citep[cf.][]{stc99}." }, "0403/astro-ph0403685_arXiv.txt": { "abstract": "More than half of all low-redshift AGN exhibit UV and X-ray absorption by highly ionized gas. The observed UV and X-ray absorption lines are almost always blue-shifted at velocities of hundreds of km/s, indicating that the absorbing gas is outflowing from the active nucleus. In some cases the inferred mass flux rivals the Eddington limit of the central black hole, an indication that these outflows are intimately related to the mass accretion and energy generation mechanism in AGN. The ejected material can also have an affect on the interstellar medium of the host galaxy and the surrounding intergalactic medium. Over the past several years, coordinated UV and X-ray observations of several bright AGN at high spectral resolution using {\\it HST}, {\\it FUSE}, {\\it Chandra}, and {\\it XMM-Newton} have contributed greatly to our understanding of these outflows. I will give an overview of these recent observations, summarize our {\\it FUSE} survey of low-redshift AGN, and interpret the results in the context of models of winds from accretion disks and thermally driven winds from the obscuring torus. ", "introduction": "Mass outflows from active galactic nuclei (AGN) can profoundly affect the evolution of the central engine \\citep{BB99}, the host galaxy and its interstellar medium \\citep{SR98, WL03} and also the surrounding intergalactic medium (IGM) \\citep{Cavaliere02, Granato04, SO04}. Winds from the high metal-abundance nuclear regions may be a significant source for enriching the IGM \\citep{Adelberger03}. Absorption by the outflow can also collimate the ionizing radiation \\citep{Kriss97} and thereby influence the ionization structure of the host galaxy and the surrounding IGM. The outflowing gas in AGN is sometimes visible as extended, bi-conical emission at visible or X-ray wavelengths (e.g., NGC 4151, \\citealt{Evans93}, \\citealt{Hutchings98}; NGC 1068, \\citealt{Ogle03}), but it most frequently manifests itself as blue-shifted absorption features in their UV and X-ray spectra. About half of all low-redshift AGN show X-ray absorption by highly ionized gas \\citep{Reynolds97, George98}, and a similar fraction show associated UV absorption in ionized species such as {\\sc C~iv} \\citep{Crenshaw99} and {\\sc O~vi} \\citep{Kriss01}. In more luminous quasars, the fraction of objects in the Sloan Digital Sky Survey that shows broad {\\sc C~iv} absorption troughs rises steeply as the troughs become narrower \\citep{Tolea02,Reichard03}, comprising over 30\\% of the quasar population at widths narrower than 1000 \\kms. For AGN that have been observed in both the X-ray and the UV, there is a one-to-one correspondence between objects showing X-ray and UV absorption, implying that the phenomena are related in some way (Crenshaw et al. 1999). The high frequency of occurrence of UV and X-ray absorption suggests that the absorbing gas has a high covering fraction, and that it is present in all AGN. The gas has a total mass exceeding $\\sim 10^3~\\rm M_\\odot$ (greater than the broad-line region, or BLR), and is outflowing at a rate $>0.1~\\rm M_\\odot~yr^{-1}$ ($10 \\times$ the accretion rate in some objects) \\citep{Reynolds97}. Key questions for understanding the outflowing, absorbing gas in AGN are: \\vspace{-12pt} \\begin{list}{} % {\\listparindent 0pt \\leftmargin 0pt \\rightmargin 0pt \\itemindent -0pt\\parsep 0pt\\labelsep 0pt\\itemsep 0pt} \\begin{sloppy} \\smallskip \\item {$\\bullet~~~$} What are the column densities and ionic abundances in the absorbing gas?\\\\ Even such a basic question is uncertain since the absorption line profiles are complex. Doublet ratios show that the absorbers can be optically thick, but they are not black at line center. Thus column densities are frequently underestimated, sometimes by as much as an order of magnitude \\citep{Arav02, Arav03}. \\smallskip \\item {$\\bullet~~~$} Where is the absorbing gas located?\\\\ The location of the absorbers is a vital clue to the process producing the outflow. Winds arising from an accretion disk \\citep{KK94, Murray95, Elvis00, Proga00} will have material at a broad range of radii reaching from near the disk itself to beyond the BLR. Thermally driven winds arising from the obscuring torus \\citep{KK95, KK01} will lie at much larger radii, typically at distances of $\\sim 1$ pc. The radial location also determines the mass flux---the larger the distance, the higher the total mass and the mass flux in the wind. \\smallskip \\item {$\\bullet~~~$} Do variations reflect an ionization response, or are they due to bulk motion?\\\\ In the limited number of monitoring campaigns carried out thus far, examples of both have been seen. The neutral hydrogen and the {\\sc C~iii} absorption in NGC~4151 responds quite clearly to continuum variations \\citep{Kriss97, Espey98}. Some absorbing clouds in NGC~3783 have shown changes consistent with an ionization response, while others appear to be due to bulk motion \\citep{Crenshaw99, Gabel03b}. \\smallskip \\item {$\\bullet~~~$} How are the X-ray and UV absorption related?\\\\ In some cases, UV absorbing gas may be directly associated with the X-ray warm absorber (3C351: \\citealt{Mathur94}; NGC~5548: \\citealt{Mathur95}; NGC 3516: \\citealt{Kraemer02}). In the extensive recent {\\it Chandra/FUSE/HST} campaign on NGC 3783 \\citep{Kaspi02, Gabel03a}, the kinematics of the X-ray absorbing gas are also a good match to the UV-absorbing gas. In other cases, however, the UV gas appears to be in an even lower ionization state, and there is no direct relation between the X-ray absorption and the multiple kinematic components seen in the UV (NGC~4151: \\citealt{Kriss95}, \\citealt{Kraemer01}; NGC~3516: \\citealt{Kriss96a, Kriss96b}; NGC~5548: \\citealt{Mathur99}, \\citealt{Crenshaw03}; Mrk~509: \\citealt{Kriss00b}, \\citealt{Yaqoob03}; NGC~7469: \\citealt{Kriss00a, Kriss03, Blustin03}). The X-ray absorbing gas itself contains material spanning a large range of ionization parameters \\citep{Lee02, Sako03, Netzer03}, and it is likely that this broad range of physical conditions can also include the UV-absorbing ions. This is a natural prediction of the thermally driven wind model of \\citet{KK95, KK01}, and would also be likely in disk-driven winds. \\end{sloppy} \\end{list} \\vspace{-8pt} To address some of these questions, we have been conducting a survey of the $\\sim100$ brightest AGN using the {\\it Far Ultraviolet Spectroscopic Explorer (FUSE)}. The short wavelength response (912--1187 \\AA) of {\\it FUSE} \\citep{Moos00} enables us to make high-resolution spectral measurements ($R \\sim 20,000$) of the high-ionization ion {\\sc O~vi} and the high-order Lyman lines of neutral hydrogen. As of November 1, 2002, we have observed a total of 87 AGN. Of these, 57 have $z < 0.15$, so that the {\\sc O~vi} doublet is visible in the {\\it FUSE} band. The \\ovi\\ doublet is a crucial link between the higher ionization absorption edges seen in the X-ray and the lower ionization absorption lines seen in earlier UV observations. The high-order Lyman lines provide a better constraint on the total neutral hydrogen column density than Ly$\\alpha$ alone. Lower ionization species such as {\\sc C~iii} and {\\sc N~iii} also have strong resonance lines in the \\FUSE\\ band, and these often are useful for setting constraints on the ionization level of any detected absorption. The Lyman and Werner bands of molecular hydrogen also fall in the \\FUSE\\ band, and we have searched for (but not found) intrinsic $\\rm H_2$ absorption that may be associated with the obscuring torus. \\begin{figure}[h] \\plottwo{gkrissf1a.eps}{gkrissf1b.eps} \\caption{ Left: Histogram of FUSE AGN versus luminosity. The shaded area shows the number of objects exhibiting intrinsic absorption. Right: The points show outflow velocity as a function of luminosity. \\label{fig1}} \\end{figure} ", "conclusions": "The multiple kinematic components frequently seen in the UV absorption spectra of AGN clearly show that the absorbing medium is complex, with separate UV and X-ray dominant zones. In some cases, the UV absorption component corresponding to the X-ray warm absorber can be clearly identified (e.g., Mrk~509, \\citealt{Kriss00b}). In others, however, {\\it no} UV absorption component shows physical conditions characteristic of those seen in the X-ray absorber (NGC~3516, \\citealt{Kriss96a, Kriss96b}; NGC~5548, \\citealt{Brotherton02}). One potential geometry for this complex absorbing structure is high-density, low-column UV-absorbing clouds embedded in a low-density, high-ionization medium that dominates the X-ray absorption. Disk-driven winds are a possible explanation for some cases of AGN outflows. By analogy to stellar winds, one would expect the terminal velocity of an AGN outflow to reflect the gravity of its origin. Disk-driven winds should therefore have velocities in the range of several thousand \\kms. Objects with broad, smooth profiles might fall in this category. The geometry proposed by \\citet{Elvis00} suggests that these objects should have only modest inclinations. However, two prime examples of Seyferts with broad smooth absorption troughs, NGC 3516 \\citep{Hutchings01} and NGC 4151 \\citep{Kriss01}, are likely the highest inclination sources in our sample given their extended, bi-conical narrow emission-line region morphologies \\citep{Miyaji92, Evans93} and their opaque Lyman limits \\citep{Kriss97}. The lower velocities we observe in objects like NGC 3783 and NGC 5548 are more compatible with thermally driven winds from the obscuring torus \\citep{KK95, KK01}. In these thermally driven winds, photoionized evaporation in the presence of a copious mass source (the torus) locks the ratio of ionizing intensity to total gas pressure (the ionization parameter $\\Xi$) at a critical value. For AGN spectral energy distributions lacking a strong extreme ultraviolet bump, such as the composite spectra of quasars assembled by \\citet{Zheng97}, \\citet{Laor97}, and \\citet{Telfer02}, the ionization equilibrium curve exhibits an extensive vertical branch. Thus, at the critical ionization parameter for evaporation, there is a broad range of temperatures that can coexist in equilibrium at nearly constant pressure. For this reason, the flow is expected to be strongly inhomogeneous. Outflow velocities are typical of the sound speed in the heated gas, or several hundred \\kms, comparable to the velocities seen in many AGN. In summary, we find that \\ovi\\ absorption is common in low-redshift ($z < 0.15$) AGN. 30 of 53 Type 1 AGN with $z < 0.15$ observed using \\FUSE\\ show multiple, blended \\ovi\\ absorption lines with typical widths of $\\sim 100~\\kms$ that are blueshifted over a velocity range of $\\sim$ 1000 \\kms. Those galaxies in our sample with existing X-ray or longer wavelength UV observations also show {\\sc C~iv} absorption and evidence of a soft X-ray warm absorber. In some cases, a UV absorption component has physical properties similar to the X-ray absorbing gas, but in others there is no clear physical correspondence between the UV and X-ray absorbing components." }, "0403/astro-ph0403016_arXiv.txt": { "abstract": "X-ray and Sunyaev-Zeldovich Effect data can be combined to determine the distance to galaxy clusters. High-resolution X-ray data are now available from the {\\it Chandra} Observatory, which provides both spatial and spectral information, and Sunyaev-Zeldovich Effect data were obtained from the BIMA and OVRO arrays. We introduce a Markov chain Monte Carlo procedure for the joint analysis of X-ray and Sunyaev-Zeldovich Effect data. The advantages of this method are the high computational efficiency and the ability to measure simultaneously the probability distribution of all parameters of interest, such as the spatial and spectral properties of the cluster gas and also for derivative quantities such as the distance to the cluster. We demonstrate this technique by applying it to the {\\it Chandra} X-ray data and the OVRO radio data for the galaxy cluster Abell~611. Comparisons with traditional likelihood-ratio methods reveal the robustness of the method. This method will be used in follow-up papers to determine the distances to a large sample of galaxy clusters. ", "introduction": "Analysis of Sunyaev-Zeldovich Effect (SZE) and X-ray data provides a unique method of directly determining distances of galaxy clusters. Clusters of galaxies contain hot plasma ($k_B T_e \\sim$ 2-20 keV) that scatters the cosmic microwave background radiation (CMB). On average, this inverse Compton scattering boosts the energy of the CMB photons, causing a small distortion in the CMB spectrum (Sunyaev and Zeldovich 1970,1972). For recent reviews of the SZE and its application for cosmology see Birkinshaw (1999) and Carlstrom et al. (2002). The SZE is proportional to the integrated pressure along the line of sight, $\\Delta T \\propto \\int n_e T_e dl$, where $n_e$ and $T_e$ are the electron density and electron temperature of the cluster plasma. The thermal X-ray emission from the same plasma has a different dependence on the density, $S_x \\propto \\int n_e^2 \\Lambda_{ee} dl$, where $\\Lambda_{ee}$ is the X-ray cooling function. Making assumptions on the distribution of the plasma (e.g., a $\\beta$ profile) and taking advantage of the different dependence on $n_e$, SZE and X-ray observations can be combined to determine the distance to galaxy clusters. These cluster distances can be combined with redshift measurements to determine the value of the Hubble constant (Myers et al. 1997, Grainge et al. 2002, Reese et al. 2002). In this paper we introduce a Markov chain Monte Carlo (MCMC) procedure for the joint analysis of SZE and X-ray data. The method is tested on the galaxy cluster Abell 611 which has SZE data obtained with the Caltech millimeter interferometric array at the Owens Valley Radio Observatory (OVRO) outfitted with centimeter-wave receivers (Carlstrom, Joy and Grego 1996) and X-ray data from the {\\it Chandra} X-ray Observatory. In a subsequent paper we will report the application of this technique to a sample of $\\sim$40 clusters. ", "conclusions": "We present a Markov chain Monte Carlo technique to derive cluster distances from SZE and X-ray data. The method was succesfully tested on the OVRO and {\\it Chandra} data of Abell~611, a galaxy cluster at z=0.288. We measure an angular diameter distance of $D_A=1.00\\pm^{0.24}_{0.21}$ Gpc (68\\% confidence level). In a previous work based on the same SZE data, but using lower-resolution X-ray data, Reese et al. (2002) derived a distance of $D_A=0.99\\pm^{0.32}_{0.29}$ Mpc. The {\\it Chandra} X-ray data provides simultaneous spatial and spectral information, featuring the finest angular resolution to date ($\\sim 0.5$ arcsec half-power radius). The MCMC method has two major advantages: it is computationally more efficient than the traditional likelihood ratio-based methods and it provides simultaneously the probability distribution function of all model parameters. This technique will be used in future papers to determine the distances of a large sample of galaxy clusters for which there are available high-resolution {\\it Chandra} X-ray data and BIMA/OVRO SZE data. This work is supported by NASA LTSA grant NAG 5-7986. E.D.R. acknowledges support from NASA Chandra Postdoctoral Fellowship PF 1-20020. Partial support was also provided by NSF grants AST-0096913 and PHY-0114422. S.J.L.\\ acknowledges support from NASA GSRP Fellowship NGT5-50173. We thank the referee for helpful comments and suggestions. \\newpage" }, "0403/astro-ph0403220_arXiv.txt": { "abstract": "We use the lag-luminosity relation to calculate self-consistently the redshifts, apparent peak bolometric luminosities $L_B$, and isotropic energies $E_{\\rm iso}$ for a large sample of BATSE gamma-ray bursts. We consider two different forms of the lag-luminosity relation; for both forms the median redshift for our burst database is 1.6. We model the resulting $E_{\\rm iso}$ sample with power law and Gaussian probability distributions without redshift evolution, both of which are reasonable models. The power law model has an index of $\\alpha_E=1.76\\pm0.05$ (95\\% confidence), where $p(E_{\\rm iso}) \\propto E_{\\rm iso}^{-\\alpha_E}$. The simple universal jet profile model suggested but did not require $\\alpha_E=2$, and subsequent physically reasonable refinements to this model permit greater diversity in $\\alpha_E$, as well as deviations from a power law; therefore our observed $E_{\\rm iso}$ probability distribution does not disprove the universal jet model. ", "introduction": "The major breakthroughs in the study of gamma-ray bursts of the past six years---most if not all bursts are cosmological, bursts do {\\it not} have constant peak luminosity, the fireballs are beamed, many bursts are associated with supernovae---resulted from the intensive study of a relatively small number of bursts without regard for whether these bursts formed a well defined statistical sample. However, realizing that for any one burst we are only sampling an anisotropic radiation pattern from one direction, we now have to study distributions of burst properties to reconstruct the bursts' appearance from all directions. Well-defined burst samples are required to derive these distributions. Unfortunately, we do not yet have a large sample of bursts with spectroscopic redshifts, which in most cases are required for calculating the intrinsic burst properties. The two dozen or so bursts with redshifts were detected by various detectors with different detection thresholds and energy sensitivities, and the follow-ups that determined the redshifts depended on the vagaries of weather and telescope availability. BATSE provided a large burst sample with well-understood thresholds, but without direct redshift determinations for most of these bursts. However, redshifts can be determined indirectly from the lag-luminosity (Norris, Marani \\& Bonnell 2000) or variability-luminosity (Fenimore \\& Ramirez-Ruiz 2000; Reichart et al. 2001) relations. Here we calculate self-consistently the redshifts for a large subset of the BATSE bursts using the lag-luminosity relation. To accommodate the problematic burst GRB980425, which is significantly underluminous if the lag-luminosity relation is a single power law, Salmonson (2001) and Norris (2002) proposed that the relation should be a broken power law. This reduces the luminosity of many long-lag bursts, moving this population closer to the observer. We use the resulting redshifts to calculate both the peak bolometric luminosity $L_B$ and the isotropic energy $E_{\\rm iso}$. Bursts are thought to radiate anisotropically, but we sample their radiation field in only one direction. Therefore $L_B$ and $E_{\\rm iso}$ are calculated from the observed flux as if the emissions are isotropic; corrections for the anisotropy are based on models of the relativistic jets that emit the observed gamma rays (see Bloom, Frail \\& Kulkarni 2003). $L_B$ and $E_{\\rm iso}$ are both bolometric quantities in the burst frame. $L_B$ is the maximum value of the luminosity while $E_{\\rm iso}$ is the total energy emitted over the duration of the burst (without correcting for the anisotropic emission). We consider the isotropic energy $E_{\\rm iso}$ to be more fundamental, and therefore model its probability distribution. Two models have been proposed for the structure of the jets. The uniform jet model (Frail et al. 2001; Bloom et al. 2003) assumes that all jets have a constant surface energy density $\\epsilon$ (energy per solid angle across the jet) but differing opening angles $\\theta_0$ (the angle between the jet axis and the edge of the jet). The model makes no predictions for the energy probability distribution. On the other hand, in the universal jet profile model (Rossi, Lazzati, \\& Rees 2002; Zhang \\& Meszaros 2002) all jets have the same surface energy density $\\epsilon$ as a function of off-axis angle $\\theta$ (the angle from the jet axis), and thus the observed differences in $L_B$ or $E_{\\rm iso}$ result from the angle $\\theta_v$ between the jet axis and the line-of-sight. This model predicts that the energy probability distribution is a power law with index $\\alpha_E$, where $p(E_{\\rm iso}) \\propto E_{\\rm iso}^{-\\alpha_E}$. If $\\epsilon \\propto \\theta^k$, then $\\alpha_E=1-2/k$. Rossi et al. (2002) and Zhang \\& Meszaros (2002) suggested $k=-2$, resulting in $\\alpha_E=2$, to reproduce the distributions observed by Frail et al. (2001), although this was not a firm prediction. A Gaussian surface energy density profile results in $\\alpha_E=1$ (Lloyd-Ronning, Dai \\& Zhang 2004). Thus $\\alpha_E$ is of order 1--2. Lloyd-Ronning et al. (2004) showed that allowing the parameters of the jet profile to vary somewhat from burst to burst---the jet profile is now only quasi-universal---results both in variations in the value of $\\alpha_E$ and deviations from a pure power law. While numerical modelling of hypernovae (Zhang, Woosley, \\& MacFadyen 2003) shows that the surface energy density of the outflows should vary with off-axis angle, it does not predict the profile; thus the profiles are fitted empirically to the data. Consequently, the observed $E_{\\rm iso}$ probability distribution is not a powerful discriminant between the two jet structure models. This paper has two goals. First, we calculate the redshift of a large sample of bursts self-consistently from the lag-luminosity relation. With the redshift and the observed fluxes and fluences we calculate $L_B$ and $E_{\\rm iso}$. Second, we model the $E_{\\rm iso}$ distribution with two functional forms: power law and lognormal. We find the best fit values for the parameters of each functional form, and evaluate how well each functional form describes the data. We also compare the fit of the power law functional form to the predictions of the universal jet profile model. In these calculations we use a cosmology with $H_0 = 70$~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_m = 0.3$ and $\\Omega_\\Lambda = 0.7$. The notation $p(a\\,|\\,b)$ means the probability of $a$ given $b$. Lower case $p$ denotes a probability density while upper case $P$ (without a subscript) represents a cumulative probability. Note that $a$ and $b$ are propositions that are true or false. A proposition can be a statement such as ``the energy distribution can be described with a lognormal functional form'' or a parameter value (equivalent to the statement ``the parameter value is 5''). In \\S 2 we present the methodology used in this study: calculating the burst redshifts (\\S 2.1); fitting the energy probability distribution (\\S 2.2); and using the cumulative probability to test the quality of the fit (\\S 2.3). The implementation of this methodology is discussed in \\S 3 and the results are provided by \\S 4. Finally, our conclusions are in \\S 5. ", "conclusions": "In this paper we have two objectives. First, we use the lag-luminosity relation to calculate self-consistently the redshifts for 1218 BATSE bursts. For the bursts without the spectral parameters required by the calculation we use average low and high energy spectral indices, and a peak energy $E_p$ derived from the hardness ratio. We find that the redshift is quite sensitive to the spectral parameters, particularly if the spectrum is sharply peaked. We use both single power law and broken power law lag-luminosity relations, and find that the broken power law relation does indeed predict a population of low luminosity, nearby bursts. For both forms of the relation the median redshift is 1.58. We use the redshifts to calculate the apparent peak bolometric flux and the isotropic energy, both assuming that the bursts radiate isotropically. Second, we fit two functional forms to the distribution of the isotropic energy. We find that our burst data can be fit by a power law energy distribution with $\\alpha_E=1.76\\pm0.05$ (95\\% confidence); considering the likely systematic uncertainties in addition to the statistical variance, the power law distribution is probably a good description of the data. This value of $\\alpha_E$ is in the acceptable range for the universal jet profile model, and therefore, our work does not distinguish between the current jet structure models. A lognormal energy distribution also describes the data; the data permit a smaller average energy if the distribution is wider. More faint bursts will bound the lower end of this distribution." }, "0403/astro-ph0403546_arXiv.txt": { "abstract": " ", "introduction": "One of the key puzzles in X-ray astronomy is to understand accretion flows in a strong gravitational field. This applies to both Active Galactic Nuclei (AGN) and Quasars, where the accretion is onto a supermassive black hole, the puzzling ultraluminous compact sources which may be intermediate mass black holes, the stellar mass Galactic black holes (GBH) and even the neutron star systems. Neutron star radii are of order three Schwarzchild radii, i.e. the same as that for the last stable orbit of material around a black hole. Thus they have very similar gravitational potentials so should have very similar accretion flows, though of course with the major difference that neutron stars have a solid surface, so can have a boundary layer and a stellar magnetic field. The main premise is that progress in understanding accretion in {\\em any} of these objects should give us some pointers to understanding accretion in {\\em all} of them. Galactic sources are intrinsically less luminous, but a great deal closer than the AGN, so generally are much brighter. Galactic black holes are also often generically transient, showing large variability on timescales from milliseconds to years. These give us a sequence of spectra at differing mass accretion rates onto the central object, allowing us to test accretion models. ", "conclusions": "We can form a unified picture of the spectral evolution of all types of low mass X-ray binaries, where the major hard-soft spectral transition is driven by a changing inner disc radius linked to the collapse of an optically thin, hot inner flow. We see clear differences between the black holes and disc accreting neutron stars, both in the form of a unique black hole spectral signature (the high/soft state) and in terms of the evolution of their spectral shape with $L/L_{Edd}$. These differences can be modeled qualitatively and quantitatively as the {\\em same} accretion flow onto a {\\em different} object: neutron stars have a surface so have a boundary layer, while black holes have an event horizon!" }, "0403/astro-ph0403400_arXiv.txt": { "abstract": "Calculations of weak-interaction transition rates and of nuclear formation enthalpies show that in isolated neutron stars, the solid phase, above the neutron-drip threshold, is amorphous and heterogeneous in nuclear charge. The neutrino emissivities obtained are very dependent on the effects of proton shell structure but may be several orders of magnitude larger than the electron bremsstrahlung neutrino-pair emissivity at temperatures $\\sim 10^{9}$ K. In this phase, electrical and thermal conductivities are much smaller than for a homogeneous {\\it bcc} lattice. In particular, the reduced electrical conductivity, which is also temperature-independent, must have significant consequences for the evolution of high-multipole magnetic fields in neutron stars. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403636_arXiv.txt": { "abstract": "{We have surveyed the metal-rich spiral galaxy M~83 (NGC~5236) for its Wolf-Rayet population using VLT-FORS2 narrow-band imaging and follow-up spectroscopy. From a total of 280 candidates identified using He\\,{\\sc ii} $\\lambda$4686 imaging, Multi Object Spectroscopy of 198 sources was carried out, revealing 132 objects containing bona-fide Wolf-Rayet features. From this sample, an exceptional W-R content of $\\sim$1030 is inferred, with N(WC)/N(WN)$\\sim$1.3, continuing the trend to larger values at higher metallicity amongst Local Group galaxies. More dramatic is the dominance of late-type WC stars in M~83 with N(WC8--9)/N(WC4--7)=9 which we attribute to the sensitivity of the classification line C\\,{\\sc iii} $\\lambda$5696 to mass-loss, providing the strength of WC winds scale with metallicity. One young massive compact cluster, \\#74 in our catalogue, hosts 20\\% of the entire galactic population, namely $\\sim$180 late WC stars and $\\sim$50 late WN stars. ", "introduction": "Wolf-Rayet stars represent excellent diagnostics of recent star formation in starbursts. Though few in number, they possess very powerful winds and so contribute significantly to the energy released into the ISM in young starbursts (Crowther \\& Dessart 1998), and are believed to represent the immediate precursors of Type Ib/c Supernovae and long duration Gamma-Ray Bursts. Individual WR stars are readily identified in nearby galaxies via their characteristic strong, broad emission lines, whilst their collective presence can be seen in the integrated light of distant starbursts. Mass-loss rates of hot, massive stars prior to the WR phase are known to scale with metallicity (Vink et al. 2001), such that the relative number of WR stars increases with higher metallicity (Maeder \\& Meynet 1994). In addition, the distribution amongst nitrogen-rich WN versus carbon-rich WC subclasses is also known to vary with metallicity for Local Group galaxies (Massey \\& Johnson 1998). Studies of resolved WR populations at metallicities higher than the Solar value have proved difficult to date because the inner Milky Way is visibly obscured, whilst the other metal-rich Local Group member, M31, lies at an unfavourable inclination. Towards this goal, M~83 (NGC~5236) is a massive spiral galaxy that is believed to be particularly metal-rich: log(O/H)+12=9.0--9.3 as derived from strong-line nebular methods (Bresolin \\& Kennicutt 2002)\\footnote{However, see Kennicutt et al. (2003) regarding the reliability of strong nebular line techniques at high metallicity.} and is the focus of the present Letter. A more detailed discussion of the WR population of M~83 and presentation of the full catalogue is given by Hadfield et al. (in preparation). ", "conclusions": "\\label{sect5} We have surveyed the nearby metal-rich galaxy M~83 for the presence of WR stars using VLT-FORS2. From follow up spectroscopy of 198 out of 280 candidates, 132 regions containing WR stars are identified. Assuming intrinsic line fluxes comparable to Galactic counterparts (Schaerer \\& Vacca 1998), we identify 1030 WR stars in M~83, with N(WC)/N(WN)$\\sim$1.3, continuing the trend observed amongst Local Group galaxies to higher metallicity (Massey \\& Johnson 1998). Accounting for the remaining candidates the total WR population may be as high as 1500. More than 50\\% of the known WR are identified as late WC stars, versus $\\leq$5\\% early WC stars, which is unprecedented relative to more metal-poor Local Group galaxies. One young massive compact cluster, \\#74, hosts $\\sim$230 late WN and WC stars. The relatively large WC population with increasing metallicity can readily be understood from comparison with evolutionary models (Maeder \\& Meynet 1994). At higher metallicity, mass-loss rates during and subsequent to the main-sequence evolution of massive stars strips away higher layers earlier on, such that a star with a particular initial mass advance to later (WC) phases at higher metallicity. But why are there exclusively {\\it early } WC stars in the LMC, a mixed population in the Milky Way and almost exclusively {\\it late} WC stars in M~83? A decade ago it was believed that (C+O)/He increases from WC9 to WC4 (Smith \\& Maeder 1991). However, subsequent spectral analysis failed to confirm any systematic trend in C/He with subtype (Koesterke \\& Hamann 1995; Crowther et al. 2002). If carbon abundances do not play a dominant role, what does? We suggest differences in wind densities are primarily responsible. The wind densities of WC stars in the LMC are $\\sim$50\\% lower than those of their Milky Way counterparts according to Crowther et al. (2002), which they attributed to a metallicity dependence of their winds. Since C\\,{\\sc iii} $\\lambda$5696 increases dramatically in strength with increasing mass-loss rate (see their Fig.12), one would expect yet higher mass-loss rates and even stronger C\\,{\\sc iii} $\\lambda$5696 in M~83 WC stars if WC winds are metallicity dependent. Our present results are fully consistent with such a population. A genuine metallicity dependence of WR winds has implications for the hard ionizing fluxes of young starbursts (Smith et al. 2002). Regardless, the presence of late WC stars is definitely a indicator of a metal-rich environment. Finally, should a supernova be observed in M~83 in the near future, we now possess a reasonable statistical sample with which we should be able to verify whether a WR star was a potential immediate precursor." }, "0403/gr-qc0403068.txt": { "abstract": "Post-Newtonian relativistic theory of astronomical reference frames based on Einstein's general theory of relativity was adopted by General Assembly of the International Astronomical Union in 2000. This theory is extended in the present paper by taking into account all relativistic effects caused by the presumable existence of a scalar field and parametrized by two parameters, $\\beta$ and $\\gamma$, of the parametrized post-Newtonian (PPN) formalism. We use a general class of the scalar-tensor (Brans-Dicke type) theories of gravitation to work out PPN concepts of global and local reference frames for an astronomical N-body system. The global reference frame is a standard PPN coordinate system. A local reference frame is constructed in the vicinity of a weakly self-gravitating body (a sub-system of the bodies) that is a member of the astronomical N-body system. Such local inertial frame is required for unambiguous derivation of the equations of motion of the body in the field of other members of the N-body system and for construction of adequate algorithms for data analysis of various gravitational experiments conducted in ground-based laboratories and/or on board of spacecrafts in the solar system. We assume that the bodies comprising the N-body system have weak gravitational field and move slowly. At the same time we do not impose any specific limitations on the distribution of density, velocity and the equation of state of the body's matter. Scalar-tensor equations of the gravitational field are solved by making use of the post-Newtonian approximations so that the metric tensor and the scalar field are obtained as functions of the global and local coordinates. A correspondence between the local and global coordinate frames is found by making use of asymptotic expansion matching technique. This technique allows us to find a class of the post-Newtonian coordinate transformations between the frames as well as equations of translational motion of the origin of the local frame along with the law of relativistic precession of its spatial axes. These transformations depend on the PPN parameters $\\beta$ and $\\gamma$, generalize general relativistic transformations of the IAU 2000 resolutions, and should be used in the data processing of the solar system gravitational experiments aimed to detect the presence of the scalar field. These PPN transformations are also applicable in the precise time-keeping metrology, celestial mechanics, astrometry, geodesy and navigation. We consider a multipolar post-Newtonian expansion of the gravitational and scalar fields and construct a set of internal and external gravitational multipoles depending on the parameters $\\beta$ and $\\gamma$. These PPN multipoles generalize the Thorne-Blanchet-Damour multipoles defined in harmonic coordinates of general theory of relativity. The PPN multipoles of the scalar-tensor theory of gravity are split in three classes -- {\\it active}, {\\it conformal}, and {\\it scalar} multipoles. Only two of them are algebraically independent and we chose to work with the {\\it conformal} and {\\it active} multipoles. We derive the laws of conservations of the multipole moments and show that they must be formulated in terms of the {\\it conformal} multipoles. We focus then on the law of conservation of body's linear momentum which is defined as a time derivative of the {\\it conformal} dipole moment of the body in the local coordinates. We prove that the local force violating the law of conservation of the body's linear momentum depends exclusively on the {\\it active} multipole moments of the body along with a few other terms which depend on the internal structure of the body and are responsible for the violation of the strong principle of equivalence (the Nordtvedt effect). The PPN translational equations of motion of extended bodies in the global coordinate frame and with all gravitational multipoles taken into account are derived from the law of conservation of the body's linear momentum supplemented by the law of motion of the origin of the local frame derived from the matching procedure. We use these equations to analyze translational motion of spherically-symmetric and rigidly rotating bodies having finite size. Spherical symmetry is defined in the local frame of each body through a set of conditions imposed on the shape of the body and the distribution of its internal density, pressure and velocity field. We prove that our formalism brings about the parametrized post-Newtonian EIH equations of motion of the bodies if the finite-size effects are neglected. Analysis of the finite-size effects reveal that they are proportional to the parameter $\\beta$ coupled with the second and higher-order rotational moments of inertia of the bodies. The finite-size effects in the translational equations of motion can be appreciably large at the latest stage of coalescence of binary neutron stars and can be important in calculations of gravitational waveform templates for the gravitational-wave interferometers. The PPN rotational equations of motion for each extended body possessing an arbitrary multipolar structure of its gravitational field, have been derived in body's local coordinates. Spin of the body is defined phenomenologically in accordance with the post-Newtonian law of conservation of angular momentum of an isolated system. Torque consists of a general relativistic part and the PPN contribution due to the presence of the scalar field. The PPN scalar-field-dependent part is proportional to the difference between {\\it active} and {\\it conformal} dipole moments of the body which disappears in general relativity. Finite-size effects in rotational equations of motion can be a matter of interest for calculating gravitational wave radiation from coalescing binaries. ", "introduction": "\\subsection{General Conventions} Greek indices $\\alpha, \\beta, \\gamma,...$ run from 0 to 3 and mark space-time components of four-dimensional objects. Roman indices $i,j,k,...$ run from 1 to 3 and denote components of three-dimensional objects (zero component belongs to time). Repeated indices mean the Einstein summation rule, for instance, $A^\\alpha B_\\alpha=A^0B_0+A^1B_1+A^2B_2+A^3B_3$ and $T^k_k=T^1_{\\;1}+T^2_{\\;2}+T^3_{\\;3}$, etc. Minkowski metric is denoted $\\eta_{\\alpha\\beta}={\\rm diag}(-1,+1,+1,+1)$. Kronecker symbol (the unit matrix) is denoted $\\delta_{ij}= {\\rm diag}(1,1,1)$. Levi-Civita fully antisymmetric symbol is $\\varepsilon_{ijk}$ such that $\\varepsilon_{123}=+1$. Kronecker symbol is used to rise and lower Roman indices. Complete metric tensor $g_{\\alpha\\beta}$ is used to rise and lower the Greek indices in exact tensor equations whereas the Minkowski metric $\\eta_{\\alpha\\beta}$ is employed for rising and lowering indices in the post-Newtonian approximations Parentheses surrounding a group of Roman indices mean symmetrization, for example, $A_{(ij)}=(1/2)(A_{ij}+A_{ji})$. Brackets around two Roman indices denote antisymmetrization, that is $A_{[ij]}=(1/2)(A_{ij}-A_{ji})$. Angle brackets surrounding a group of Roman indices denote the symmetric trace-free (STF) part of the corresponding three-dimensional object, for instance, $$A_{}=A_{(ij)}-\\frac{1}{3}\\delta_{ij}A_{kk}\\;,\\quad\\quad A_{}=A_{(ijk)} -\\frac{1}{5}\\delta_{ij}A_{kpp}-\\frac{1}{5}\\delta_{jk}A_{ipp}-\\frac{1}{5}\\delta_{ik}A_{jpp} \\;.$$ We also use multi-index notations, for example, $$A_L=A_{i_1i_2...i_l}\\;, \\qquad\\quad B_{P-1}=B_{i_1i_2...i_{p-1}}\\;, \\qquad\\qquad D_{}=D_{}\\;.$$ Sum over multi-indices is understood as $$A_L Q^L=A_{i_1i_2...i_l}Q^{i_1i_2...i_l}\\;,\\qquad\\quad P_{aL-1} T^{bL-1}=P_{ai_1i_2...i_{l-1}}T^{bi_1i_2...i_{l-1}}\\;.$$ Comma denotes a partial derivative, for example, $\\phi_{,\\alpha}=\\partial\\phi/\\partial x^\\alpha$, where $\\phi_{,0}=c^{-1}\\partial\\phi/\\partial t,\\, \\phi_{,i}=\\partial\\phi/\\partial x^i,$ and semicolon $T^\\alpha_{\\,\\;;\\beta}$ denotes a covariant derivative. $L$-order partial derivative with respect to spatial coordinates is denoted $\\partial_L=\\partial_{i_1}...\\partial_{i_l}$. Other conventions are introduced as they appear in the text. We summarize these particular conventions and notations in the next section for the convenience of the readers. \\subsection{Particular Conventions and Symbols Used in the Paper} \\renewcommand{\\arraystretch}{1.5} \\setlength{\\arrayrulewidth}{0.1mm} %\\setlength{\\arraycolsep}{3mm} \\begin{longtable}{|c|@{\\,}p{7cm}@{\\,}|c|} \\hline \\textbf{\\color{blue}Symbol}&\\multicolumn{1}{c|}{\\textbf{\\color{blue}Description}}& \\multicolumn{1}{c|}{\\textbf{\\color{blue}Equation(s)}} \\endhead\\hline $g_{\\mu\\nu} $&physical (Jordan-Fierz frame) metric tensor&(\\ref{10.3})\\\\\\hline $\\tilde g_{\\mu\\nu} $&conformal (Einstein frame) metric tensor&(\\ref{13.18})\\\\\\hline $ g $&the determinant of $g_{\\mu\\nu}$ &(\\ref{10.1})\\\\\\hline $\\tilde g $&the determinant of $\\tilde g_{\\mu\\nu}$&(\\ref{13.20})\\\\\\hline $\\eta_{\\mu\\nu} $&the Minkowski (flat) metric tensor&(\\ref{exp})\\\\\\hline $\\Gamma^\\alpha_{\\mu\\nu} $& the Christoffel symbol &(\\ref{covd})\\\\\\hline $R_{\\mu\\nu} $& the Ricci tensor &(\\ref{10.2})\\\\\\hline $R $& the Ricci scalar & (\\ref{10.1})\\\\\\hline $\\tilde R_{\\mu\\nu}$& the conformal Ricci tensor &(\\ref{13.19})\\\\\\hline $T_{\\mu\\nu}$& the energy-momentum tensor of matter &(\\ref{10.2})\\\\\\hline $T $& the trace of the energy-momentum tensor &(\\ref{10.2})\\\\\\hline $\\phi$& the scalar field &(\\ref{10.1})\\\\\\hline $\\phi_0 $& the background value of the scalar field $\\phi$& (\\ref{aa})\\\\\\hline $\\zeta $& the dimensionless perturbation of the scalar field &(\\ref{aa})\\\\\\hline $\\theta(\\phi) $& the coupling function of the scalar field &(\\ref{10.1})\\\\\\hline ${\\dAl}_g $& the Laplace-Beltrami operator &(\\ref{covd})\\\\\\hline $\\dAl $&the D'Alembert operator in the Minkowski space-time &(\\ref{11.29})\\\\\\hline $\\rho $& the density of matter in the co-moving frame &(\\ref{11.1})\\\\\\hline $\\rho^* $& the invariant (Fock) density of matter &(\\ref{11.19})\\\\\\hline $\\Pi $& the internal energy of matter in the co-moving frame&(\\ref{11.1})\\\\\\hline $\\pi^{\\mu\\nu} $& the tensor of (anisotropic) stresses of matter &(\\ref{11.1})\\\\\\hline $u^\\alpha $& the 4-velocity of matter &(\\ref{11.1})\\\\\\hline $v^i $&the 3-dimensional velocity of matter in the global frame &(\\ref{11.23})\\\\\\hline $\\omega$& the asymptotic value of the coupling function $\\theta(\\phi) $ &(\\ref{10.5})\\\\\\hline $ \\omega' $& the asymptotic value of the derivative of the coupling function $\\theta(\\phi) $ &(\\ref{10.5})\\\\\\hline $c$& the ultimate speed of general and special theories of relativity&(\\ref{10.1})\\\\\\hline $\\epsilon $& a small dimensional parameter, $\\epsilon=1/c$ &(\\ref{exp})\\\\\\hline $h_{\\mu\\nu} $& the metric tensor perturbation, $g_{\\mu\\nu}-\\eta_{\\mu\\nu}$ &(\\ref{mtp})\\\\\\hline $\\nne{n}{\\mu}{\\nu} $& the metric tensor perturbation of order $\\epsilon^n$ in the post-Newtonian expansion of the metric tensor &(\\ref{exp})\\\\\\hline $N $&a shorthand notation for $\\nne{2}{0}{0}$ &(\\ref{not})\\\\\\hline $L $&a shorthand notation for $\\nne{4}{0}{0}$ &(\\ref{not})\\\\\\hline $N_i $&a shorthand notation for $\\nne{1}{0}{i}$ &(\\ref{not})\\\\\\hline $L_i $&a shorthand notation for $\\nne{3}{0}{i}$ &(\\ref{not})\\\\\\hline $H_{ij} $&a shorthand notation for $\\nne{2}{i}{j}$ &(\\ref{not})\\\\\\hline $H $&a shorthand notation for $\\nne{2}{k}{k}$ &(\\ref{not})\\\\\\hline $\\tilde{N},\\; \\tilde{L}$& shorthand notations for perturbations of conformal metric $\\tilde g_{\\mu\\nu}$ &(\\ref{mp4})\\\\\\hline $\\gamma$&the `space-curvature' PPN parameter &(\\ref{11.27})\\\\\\hline $\\beta $&the `non-linearity' PPN parameter &(\\ref{11.28})\\\\\\hline $\\eta $& the Nordtvedt parameter, $\\eta =4\\beta-\\gamma-3$ &(\\ref{mp3})\\\\\\hline $G $&the observed value of the universal gravitational constant &(\\ref{10.6})\\\\\\hline ${\\cal G} $&the bare value of the universal gravitational constant &(\\ref{10.7}), (\\ref{10.6})\\\\\\hline $x^\\alpha=(x^0,x^i) $&the global coordinates with $x^0=ct$ and $x^i\\equiv{\\bm x}$ & \\\\\\hline $w^\\alpha=(w^0,w^i) $&the local coordinates with $w^0=cu$ and $w^i\\equiv{\\bm w}$ & \\\\\\hline $U $&the Newtonian gravitational potential in the global frame &(\\ref{12.5})\\\\\\hline $U^{\\sss(A)} $&the Newtonian gravitational potential of body A in the global frame &(\\ref{12.9a})\\\\\\hline $U_i $&a vector potential in the global frame &(\\ref{12.8})\\\\\\hline $U_i^{\\sss(A)} $&a vector potential of body A in the global frame &(\\ref{12.9a})\\\\\\hline $\\chi,\\;\\Phi_1,\\ldots,\\Phi_4 $&various special gravitational potentials in the global frame &(\\ref{12.6}), (\\ref{12.9ex}) \\\\\\hline $V,\\; V^i $&potentials of the physical metric in the global frame &(\\ref{mp1}), \\ref{mp2})\\\\\\hline $\\sigma,\\;\\sigma^i $&the {\\it active} mass and current-mass densities in the global frame &(\\ref{13.0}), (\\ref{13.1})\\\\ \\hline $I_{\\sss{}} $&the {\\it active} Thorne-Blanchet-Damour mass multipole moments in the global frame &(\\ref{13.9})\\\\ \\hline $S_{\\sss{}} $&the {\\it active} spin multipole moments in the global frame &(\\ref{13.10})\\\\\\hline $\\bar V $&potential of the scalar field in the global frame &(\\ref{mp3})\\\\\\hline $\\bar\\sigma $&scalar mass density in the global frame &(\\ref{13.13})\\\\\\hline $\\bar I_{\\sss{}} $&scalar mass multipole moments in the global frame &(\\ref{13.17})\\\\\\hline $\\tilde V $&gravitational potential of the conformal metric in the global frame &(\\ref{mp4})\\\\\\hline $\\tilde\\sigma $&the {\\it conformal} mass density in the global frame &(\\ref{13.21})\\\\\\hline $\\tilde I_{\\sss{}} $&the {\\it conformal} mass multipole moments in the global frame &(\\ref{13.30})\\\\\\hline $\\mathbb{M} $&conserved mass of an isolated system &(\\ref{13.35})\\\\\\hline $\\mathbb{P}^i $&conserved linear momentum of an isolated system &(\\ref{13.36})\\\\\\hline $\\mathbb{S}^i $&conserved angular momentum of an isolated system &(\\ref{13.361})\\\\\\hline $\\mathbb{D}^i $&integral of the center of mass of an isolated system &(\\ref{13.35a})\\\\\\hline $\\hat{A} $&symbols with the hat stand for quantities in the local frame & \\\\\\hline $(B) $&sub-index referring to the body and standing for the internal solution in the local frame &(\\ref{1.1}), (\\ref{1.2})\\\\\\hline $(E) $&sub-index referring to the external with respect to (B) bodies and standing for the external solution in the local frame&(\\ref{1.1}), (\\ref{1.2})\\\\\\hline $(C) $&sub-index standing for the coupling part of the solution in the local frame &(\\ref{1.2})\\\\\\hline $P_{\\sss L} $&external STF multipole moments of the scalar field &(\\ref{1.7a})\\\\\\hline $Q_{\\sss L} $&external STF gravitoelectric multipole moments of the metric tensor &(\\ref{1.8a})\\\\\\hline $C_{\\sss L} $&external STF gravitomagnetic multipole moments of the metric tensor &(\\ref{1.9a})\\\\\\hline $Z_{\\sss L},\\; S_{\\sss L} $&other sets of STF multipole moments entering the general solution for the space-time part of the external local metric&(\\ref{1.9a})\\\\\\hline $Y_{\\sss L}, B_{\\sss L}, D_{\\sss L}, E_{\\sss L}, F_{\\sss L}, G_{\\sss L}$& STF multipole moments entering the general solution for the space-space part of the external local metric &(\\ref{1.10a}) \\\\\\hline ${\\mathcal V}_i,\\;\\Omega_i $&linear and angular velocities of kinematic motion of the local frame; we put them to zero throughout the rest of the paper &(\\ref{1.8a}), (\\ref{1.8aaa})\\\\\\hline $\\nu^i $&3-dimensional velocity of matter in the local frame &(\\ref{1.12})\\\\\\hline ${\\cal I}_{\\sss{L}} $&active Thorne-Blanchet-Damour STF mass multipole moments of the body in the local frame &(\\ref{1.31})\\\\\\hline $\\sigma_{\\sss B}$&active mass density of body B in the local frame &(\\ref{pz3})\\\\\\hline $\\bar{\\cal I}_{\\sss{L}} $&scalar STF mass multipole moments of the body in the local frame &(\\ref{1.33})\\\\\\hline $\\bar\\sigma_{\\sss B} $&scalar mass density of body B in the local frame &(\\ref{pz4})\\\\\\hline $\\tilde{\\cal I}_{\\sss{L}} $&conformal STF mass multipole moments of the body in the local frame &(\\ref{1.34})\\\\\\hline $\\tilde\\sigma_{\\sss B} $&conformal mass density of body B in the local frame &(\\ref{pz5})\\\\\\hline $\\sigma^i_{\\sss B} $¤t mass density of body B in the local frame &(\\ref{pz6})\\\\\\hline $S_{\\sss L} $&spin STF multipole moments of the body in the local frame &(\\ref{1.32})\\\\\\hline $\\xi^0,\\;\\xi^i $&Relativistic corrections in the post-Newtonian transformation of time and space coordinates &(\\ref{2.2}), (\\ref{2.3})\\\\\\hline $x^i_{{\\sss B}},\\;\\nnv{i},\\;\\nnaa{i} $&position, velocity and acceleration of the body's center of mass with respect to the global frame &(\\ref{2.3}), (\\ref{2.8}), (\\ref{2.10})\\\\\\hline $\\nnr{i} $&$x^i-x^i_{{\\sss B}}(t),$ i.e. the spacial coordinates taken with respect to the center of mass of body B in the global frame &(\\ref{2.3})\\\\\\hline ${\\cal A},\\;{\\cal B}_{\\sss{}} $&functions appearing in the relativistic transformation of time &(\\ref{2.8}), (\\ref{eq1}) \\\\\\hline ${\\cal D}_{\\sss{}},\\;{\\cal F}_{\\sss{}},\\;{\\cal E}_{\\sss{}} $& functions appearing in the relativistic transformation of spacial coordinates &(\\ref{eq2})\\\\\\hline $\\Lambda^{\\beta}_{\\;\\,\\alpha} $&matrix of transformation between local and global coordinate bases &(\\ref{cb})\\\\\\hline $\\mathfrak{B},\\;\\mathfrak{D},\\;\\mathfrak{B}^i,\\;\\mathfrak{P}^i,\\;\\mathfrak{R}^{i}_{\\;j} $& PN corrections in the matrix of transformation $\\Lambda^{\\beta}_{\\;\\,\\alpha}$&(\\ref{ma00})--(\\ref{ma03})\\\\\\hline $\\bar U,\\;\\bar U^i,\\;$ etc.& external gravitational potentials &(\\ref{3.1})--(\\ref{3.4})\\\\\\hline $\\bar U_{,L}\\nnxe,\\;\\bar U^i_{,L}\\nnxe $& $l$-th spatial derivative of an external potential taken at the center of mass of body B &(\\ref{3.13})\\\\\\hline $\\mathcal{U}^{{\\sss(B)}} $& PN correction in the formula of matching of the local Newtonian potential &(\\ref{3.5})\\\\\\hline ${F}^{ik} $&the matrix of relativistic precession of local coordinates with respect to global coordinates&(\\ref{5.8})\\\\ \\hline ${\\cal M}_*,\\;{\\cal J}^i_*,\\;{\\cal P}^i_* $& Newtonian-type mass, center of mass, and linear momentum of the body in the local frame &(\\ref{a})--(\\ref{c})\\\\\\hline ${\\rm M} $& general relativistic PN mass of the body in the local frame &(\\ref{ij3})\\\\\\hline ${\\cal M} $& active mass of the body in the local frame &(\\ref{ij2})\\\\\\hline $\\tilde {\\cal M} $& conformal mass of the body in the local frame &(\\ref{ij1})\\\\\\hline ${\\cal I}^{\\sss{(2)}} $&rotational moment of inertia of the body in the local frame &(\\ref{fop})\\\\\\hline ${\\cal N}^{\\sss L} $& a set of STF multipole moments &(\\ref{aer})\\\\\\hline ${\\cal P}^i $& PN linear momentum of the body in the local frame &(\\ref{6.2b})\\\\\\hline $\\Delta\\dot{\\cal P}^i $& scalar-tensor PN correction to $\\dot{\\cal P}^i$ &(\\ref{bvo})\\\\\\hline $\\tilde{\\cal M}_{ij} $&conformal anisotropic mass of the body in the local frame &(\\ref{brt})\\\\\\hline ${\\mathbb {F}}^i_N,\\;\\Delta{\\mathbb {F}}^i_{N},\\;{\\mathbb {F}}^i_{pN},\\;\\Delta{\\mathbb {F}}^i_{pN} $& gravitational forces in the expression for $Q_i$ &(\\ref{6.9})--(\\ref{fopm})\\\\\\hline ${\\cal S}^i$&the bare post-Newtonian definition of the angular momentum (spin) of a body&(\\ref{spin-3}) \\\\\\hline ${\\cal T}^i$&the post-Newtonian torque in equations of rotational motion&(\\ref{spin-5}) \\\\\\hline $\\Delta{\\cal T}^i$&the post-Newtonian correction to the torque ${\\cal T}^i$ &(\\ref{spin-6}) \\\\\\hline $\\Delta{\\cal S}^i$&the post-Newtonian correction to the bare spin ${\\cal S}^i$ &(\\ref{spin-7}) \\\\\\hline ${\\cal R}^i$&velocity-dependent multipole moments &(\\ref{spin-8}) \\\\\\hline ${\\cal S}^i_+$&the (measured) post-Newtonian spin of the body&(\\ref{spin-9})\\\\\\hline $r $& radial space coordinate in the body's local frame, $r=|{\\bm w}|$ &(\\ref{as7})\\\\\\hline ${\\Omega}^j_{\\sss B} $& angular velocity of rigid rotation of the body B referred to its local frame &(\\ref{qma})\\\\\\hline $I^{(2l)}_{\\sss B} $& $l$-th rotational moment of inertia of the body B &(\\ref{9.6.3})\\\\\\hline $\\mathbb{I}_{\\sss B}^L $& multipole moments of the multipolar expansion of the Newtonian potential in the global coordinates &(\\ref{aq9})\\\\\\hline $R_{\\sss B} $&$|{\\bm R}_{\\sss B}|$, where ${\\bm R}_{\\sss B}={\\bm x}-{\\bm x}_{\\sss B}$ &(\\ref{re4})\\\\\\hline $R_{\\sss{BC}}^i $&$x^i_{\\sss C} - x^i_{\\sss B}$ &(\\ref{pn7})\\\\\\hline $F^i_{\\sss N},\\;F^i_{\\sss{EIH}},\\;F^i_{\\sss{\\cal S}},\\;F^i_{\\sss{\\cal I}GR},\\; \\delta F^i_{\\sss{\\cal I}GR} $& forces from the equation of motion of spherically-symmetric bodies &(\\ref{gko1})--(\\ref{gko5})\\\\ \\hline $\\mathfrak{M}_{\\sss B} $&Nordtvedt's gravitational mass of the body B &(\\ref{dm})\\\\\\hline \\end{longtable} ", "conclusions": "" }, "0403/astro-ph0403510_arXiv.txt": { "abstract": "Optical and infrared monitoring of the afterglow site of gamma-ray burst (GRB) 031203 has revealed a brightening source embedded in the host galaxy, which we attribute to the presence of a supernova (SN) related to the GRB (``SN 031203''). We present details of the discovery and evolution of SN 031203 from 0.2 to 92 days after the GRB, derived from SMARTS consortium photometry in I and J bands. A template type Ic lightcurve, constructed from SN 1998bw photometry, is consistent with the peak brightness of SN 031203 although the lightcurves are not identical. Differential astrometry reveals that the SN, and hence the GRB, occurred less than $300 h_{71}^{-1}$ pc (3 $\\sigma$) from the apparent galaxy center. The peak of the supernova is brighter than the optical afterglow suggesting that this source is intermediate between a strong GRB and a supernova. ", "introduction": "Since the discovery of GRB afterglows, the evidence for a physical connection between gamma-ray bursts (GRBs) and core-collapse supernovae (SNe) has mounted \\citep[see reviews by][]{vanp99,mez01}. Particularly compelling were observations of lightcurves and broadband photometry of SN-like features embedded in GRB afterglow light \\citep[see][]{blo03}. Recently, spectroscopic evidence \\citep{Stanek03,Hjorth03,kdw+03} confirmed that GRBs are produced in the death of massive stars \\citep{woo93}. To date, SN signatures have been reliably found in only a few GRBs (see \\citealt{blo03}) necessitating the search for and the study of new GRB-related SNe. GRB 031203 triggered the IBIS instrument onboard the {\\it Integral} satellite on 3 December 2003 at 22:01:28 UT \\citep{Gotz03}, leading to quick discoveries of X-ray \\citep{Camp03} and radio afterglows \\citep{Frail03, SKF03}. Spectroscopy of the host galaxy coincident with the radio transient yielded a redshift of $z=0.1055$ \\citep{Proch04}, likely the redshift of the burst itself. The low redshift (second only to the unusual GRB 980425) of GRB 031203 presents a rare opportunity to create a well-sampled SN lightcurve using modest aperture telescopes. We began our observations of the field 5 hours after trigger and continued monitoring periodically for several months. We reported our discovery of an increase in brightness of the aperture magnitude of the host, and suggested the emergence of a supernova was responsible \\citep{Bailyn03}. Hereafter, since the explosion date of the SN is likely that of the GRB, we designate the SN associated with GRB 031203 as ``SN 031203''. Monitoring of the SN by other groups has now confirmed the presence of SN 031203 both photometrically \\citep{Bersier04} and spectroscopically \\citep{Tag04}. In this paper we present optical and infrared data obtained with the SMARTS 1.3m telescope and ANDICAM instrument between 0.2 and 92 days after the detection of GRB 031203. Observations and data reduction are reported in section 2. Section 3 describes the aperture photometry and image subtraction carried out on this data. The resultant evidence of a SN associated with GRB 031203 is presented. A comparison between this SN and SN 1998bw is made in section 4. ", "conclusions": "" }, "0403/astro-ph0403456_arXiv.txt": { "abstract": "We describe the design and construction of a formatted fiber field-unit, SparsePak, and characterize its optical and astrometric performance. This array is optimized for spectroscopy of low-surface brightness, extended sources in the visible and near-infrared. SparsePak contains 82, 4.7$^{\\prime\\prime}$ fibers subtending an area of 72$^{\\prime\\prime}\\times$71$^{\\prime\\prime}$ in the telescope focal plane, and feeds the WIYN Bench spectrograph. Together, these instruments are capable of achieving spectral resolutions of $\\lambda/\\Delta\\lambda\\sim 20,000$ and an area--solid-angle product of $\\sim 140$ arcsec$^2$ m$^2$ per fiber. Laboratory measurements of SparsePak lead to several important conclusions on the design of fiber termination and cable curvature to minimize focal ratio degradation. SparsePak itself has throughput $>80$\\% redwards of 5200\\AA\\, and 90-92\\% in the red. Fed at f/6.3, the cable delivers an output 90\\% encircled energy at nearly f/5.2. This has implications for performance gains if the WIYN Bench Spectrograph had a faster collimator. Our approach to integral-field spectroscopy yields an instrument which is simple and inexpensive to build, yet yields the highest area--solid-angle product per spectrum of any system in existence. An Appendix details the fabrication process in sufficient detail for others to repeat. SparsePak was funded by the National Science Foundation and the University of Wisconsin-Madison Graduate School, and is now publicly available on the WIYN Telescope through the National Optical Astronomical Observatories. ", "introduction": "Observational astronomy consists of obtaining subsets of a fundamental data hyper-cube of the apparent distribution of photons in angle$^2$ on the sky $\\times$ wavelength $\\times$ time $\\times$ polarization. Information-gathering systems (``instruments'') are designed to make science-driven trades on the range and sampling of each of these dimensions. Here we describe an instrument optimized for the study of the stellar and ionized gas kinematics in disks of nearby and distant galaxies. Such studies require bi-dimensional spectroscopy at medium spectral resolution ($5000=0.23\\pm0.01$. The luminosities of the host galaxies are found to be comparable with those of galaxies drawn from the bright end of the local cluster galaxy luminosity function, spanning the range $0.7L^{\\star}=10^{8.87\\pm 0.04}\\Msun$. Finally, a significant ($\\simeq 3\\sigma$) correlation is found between black-hole mass and 151-MHz radio luminosity for those objects in the sample with either high-excitation nuclear spectra (HEG) or classical double (CD) radio structures. ", "introduction": "The underlying mechanisms which govern the large range in low-frequency radio luminosity observed in radio-loud active galactic nuclei (AGN) are currently unknown. Over the previous two decades a wealth of observational effort has been invested in studying the redshift r\\'{e}gime $0L^{\\star}$ ellipticals (eg. Taylor et al. 1996; McLure et al. 1999; Dunlop et al. 2003). The apparent uniformity of radio-loud AGN host galaxies has taken on added importance over the last few years, following the discovery in nearby (distance $\\ltsim$ 150 Mpc) inactive galaxies that a reasonably accurate estimate ($\\Delta M_{bh}\\simeq 0.3$ dex) of the central black-hole mass can be obtained via its correlation with the mass of the host spheroidal component (Kormendy \\& Richstone 1995; Magorrian et al. 1998; Gebhardt et al. 2000; Ferrarese \\& Merritt 2000; McLure \\& Dunlop 2002; Marconi \\& Hunt 2003; Tremaine et al. 2002). Moreover, recent progress has also indicated that a similarly accurate black-hole mass estimate ($\\Delta M_{bh}\\simeq 0.4$ dex) can be obtained for broad-line AGN using emission-line widths to derive the virial mass estimate (eg. Kaspi et al. 2000; McLure \\& Dunlop 2002; McLure \\& Jarvis 2002; Vestergaard 2002). Consequently, a large body of work has appeared in the recent literature investigating the possible link between radio luminosity and black-hole mass in radio-loud AGN (eg. Laor 2000; Lacy et al. 2001; McLure \\& Dunlop 2001a; McLure \\& Dunlop 2002; Bettoni et al. 2003; Dunlop et al. 2003). Unfortunately, observational studies have traditionally been subject to a degeneracy between radio luminosity and redshift produced as a by-product of flux-limited radio samples. In order to study the properties of radio-loud AGN separated by a large dynamic range in radio luminosity, it has previously been necessary to select samples consisting of objects covering a wide range of redshifts. This has led to difficulties in interpreting the data due to the complication of potentially significant evolutionary effects. However, by selecting our sample of objects from four complete, low-frequency selected radio samples with successively fainter flux-density limits, it has been possible to construct a sample of radio galaxies which spans three decades in radio luminosity at a virtually constant cosmic epoch ($0.4110$~kK), with $\\Mdot\\sim 10^{-7}$~\\Msunyr. Our derived parameters for the [W01]-PG~1159 star Abell~78 generally agree with those of past analyses. We derive similar parameters for NGC~2371, and suggest that it is a [WR]-PG~1159 also, but its wind is more ionized and shows no \\OV\\ features (a [WO0]-PG~1159?). We find evidence of iron deficiency in both of these objects, supporting the findings of \\citet{werner:03} in Abell~78 and of \\citet{miksa:02} in PG~1159 stars. They lie on the bend in the theoretical evolutionary tracks between the constant luminosity phase and the WD cooling sequence, having post-AGB ages of 10-15~kyr. For [WC] stars, it seems that as the star evolves away from the post-AGB phase through the [WC] sequence, the temperature and terminal velocity of the wind increase as the wind density decreases \\citep{acker:03}. Except for Abell~78, the post-AGB ages predicted by the evolutionary models are typically 2-4 times lower than their kinematic ages. However, kinematic ages are lower limits to the post-AGB age (as the expansion is slower in the initial phase) thus the actual discrepancy may be smaller. Our FUV/UV analysis has provided wind parameters for H-rich and H-deficient CSPN at the stage in post-AGB evolution where the winds are fading. However, because a precise determination of the mass is not possible, the objects do not necessarily represent the same evolutionary sequence. This work has also provided information on the interstellar and circumstellar environment from our measurements of the column densities and temperatures of \\HI\\ and \\Htwo\\ along the sight-lines. Our determinations of the \\HI\\ column density and \\EBMV\\ imply that the relationship between these two quantities in the circumstellar (PN) environment of CSPN differs from that of the ISM, having lower dust-to-gas ratios (probably due to the destruction of dust by the radiation field). The high resolution of the FUSE data allow us to detect hot \\Htwo\\ associated with the nebulae. For all four objects in our sample, a single component of \\Htwo\\ gas at typical ISM temperatures (\\ie, $T \\sim 80$~kK) was not adequate to fit the absorption spectrum. A second, hotter ($T\\sim300$~kK) component was necessary, which we assume to be associated with the circumstellar environment. With the advent of many FUSE observations of CSPN, it is becoming apparent that hot circumstellar \\Htwo\\ is not uncommon CSPN at quite different evolutionary stages: from old PN with white dwarf nuclei \\citep{herald:02} to young, compact PN \\citep{herald:04a}. The objects in this paper lie either along the constant luminosity section of the post-AGB evolutionary tracks, or on the transition bend to the WD cooling sequence, and thus represent intermediate stages to the previously mentioned cases. Given the intense UV radiation fields emitted by the CSPN, it is likely that the nebular \\Htwo\\ exists in clumps, shielded by neutral and ionized hydrogen, as appears to be the case in the Helix nebula \\citep{speck:02}." }, "0403/astro-ph0403054_arXiv.txt": { "abstract": "Radio continuum observations at 20 and 6~cm of the highly inclined Virgo spiral galaxy NGC~4522 are presented. Both, 20 and 6~cm total emission distributions are asymmetric with an extended component to the west where extraplanar atomic gas and H$\\alpha$ emission are found. The 6~cm polarized emission is located at the eastern edge of the galactic disk. Its peak is located about 1~kpc to the east of the total emission peak. We argue that this phenomena is a characteristic feature for cluster galaxies which are experiencing significant pressure from the intracluster medium. The degree of polarization decreases from the east to the west. The flattest spectral index between 20 and 6~cm coincides with the peak of the 6~cm polarized emission. These findings are consistent with a picture of a large scale shock due to ram pressure located at the east of the galaxy where cosmic rays are accelerated. We conclude that it is likely that the galaxy experiences active ram pressure. ", "introduction": "The Virgo cluster spiral galaxy NGC~4522 is one of the few Virgo galaxies where we can directly observe the effects of ram pressure due to the galaxy's rapid motion in the intra-cluster medium. Kenney \\& Koopmann (1999) observed this galaxy in the optical and the H$\\alpha$ line with the WIYN telescope. Whereas the highly inclined, old stellar disk appears symmetric, the distribution of ionized gas is highly asymmetric. The H$\\alpha$ disk is sharply truncated beyond 0.35\\,$R_{25}$. Ten percent of the H$\\alpha$ emission arises from extraplanar H{\\sc ii} regions that are exclusively located to the west of the galactic disk. Kenney \\& Koopmann (1999) argue that this ionized gas distribution is reminiscent of a bow shock morphology, which suggests that the gas is pushed to the west by ram pressure. Vollmer et al. (2000) obtained an H$\\alpha$ velocity field with the Fabry-Perot Interferometer at the Observatoire de Haute Provence. The galactic disk shows a symmetrically rising rotation curve. The velocities of the extraplanar emission regions are not part of this regular rotation. Their kinematic behaviour cannot be reproduced by rotation within a gravitational potential of any known disk or halo model. Thus this gas is located out of the galactic plane and/or is accelerated/decelerated. NGC~4522 has a large line-of-sight velocity with respect to the cluster mean ($\\sim$1300~km\\,s$^{-1}$) and is located at a distance of 3.3$^{\\rm o}$ ($\\sim$1~\\footnote{We use a distance of 17~Mpc to the Virgo cluster}Mpc) south from the Virgo cluster center (M87). The intracluster medium (ICM) density at these distances might not be high enough to strip the gas at a galactic radius of 0.35\\,$R_{25}$. This is why Vollmer et al. (2000) argued that NGC~4522 has been severely affected by ram pressure in the past and the extraplanar filaments represent gas that has been pushed to larger distances from the galaxy center and is now falling back onto the galactic disk. Recently, Kenney et al. (2003) observed NGC~4522 in the H{\\sc i} 21~cm line with the VLA. They found an atomic gas distribution that is similar to that of the ionized gas: a truncated H{\\sc i} disk with two high column density blobs in the north-west and the south-west. The H{\\sc i} velocity field of the extraplanar emission regions shows clear deviations from the overall rotation pattern. Even with the distributions and velocity fields of the atomic and ionized gas it is still unclear if the galaxy is experiencing ongoing ram pressure stripping or if the extraplanar gas is falling back to the galaxy after a past ram pressure stripping event. On the one hand the galaxy's location and radial velocity excludes a simple radial orbit advocated by Vollmer et al. (2000), on the other hand with the ongoing stripping scenario it is not clear if ram pressure is strong enough at such a large distance from the cluster center where the ICM density is lower than average. Observations of the polarized radio continuum radiation can give important and complementary information on the gas dynamics as has been shown for the barred galaxy NGC~1097 (Beck et al. 1999). This radiation traces the ordered large scale ($\\sim$1~kpc) magnetic field in galactic ISM. This magnetic field is very sensitive to (i) compression and (ii) shear motion, which are both difficult to detect in radial velocity fields. Thus, enhanced polarized radio continuum radiation can be a sign of compression or strong shear motion in the gas. Both kinematic features are present during a ram pressure stripping event (see e.g. Vollmer et al. 2000): a galaxy on a highly eccentric orbit within the cluster experiences compression due to ram pressure when it passes through the cluster core (compression phase); after the core passage ram pressure ceases and the gas starts to fall back onto the galactic disk, which gives rise to strong shear motions. Observationally, there are only two Virgo cluster galaxies observed deeply in the polarized 6~cm radio continuum, because this is very time consuming (NGC~4254, Soida et al. 1996, and NGC~4654, Chyzy et al. in prep.). Both galaxies show an asymmetric distribution of polarized radio continuum emission. In order to interpret the polarization data, numerical modeling is important, because different dynamical features, such as compression or shear, can produce polarized emission maxima. Otmianowska \\& Vollmer (2003) solved the induction equation using the velocity fields of Vollmer et al. (2002) in order to calculate the evolution of the large scale magnetic field during a stripping event. In a second step they calculated the polarized radio continuum emission and made maps of its evolution that can be directly compared to observations. We will apply this method to a dynamical model designed for NGC~4522 in a forthcoming paper. In this article we present radio continuum observations at 6~cm and 20~cm obtained with the VLA. The outline of the article is as follows. The observations are described in Sec.~1. The results are discussed in Sect.~2. We compare our results to radio observation of field galaxies (Sect.~4) and compare our data with existing H{\\sc i} and H$\\alpha$ maps (Sect.~5). The discussion in Sect.~6 is followed by the conclusions in Sect.~7. ", "conclusions": "We present VLA D array observations at 6 and 20~cm of the highly inclined Virgo spiral galaxy NGC~4522. The results are \\begin{enumerate} \\item the 20~cm and 6~cm total emission distributions are asymmetric; they are more extended to the west, where extraplanar H$\\alpha$ and H{\\sc i} emission is found, \\item the 6~cm polarized emission is located at the eastern edge of the galactic disk; its peak is displaced to the east of the total emission peak, \\item the degree of polarization increases from the galaxy center towards the eastern peak and towards the northeastern edge of the disk, \\item the spectral index between 20 and 6~cm decreases from west to east and its maximum coincides with the peak of the 6~cm polarized emission, \\item polarized radio continuum emission is a powerful tool to detect interactions of spiral galaxies with their cluster environment. \\end{enumerate} Points 3 and 4 are consistent with a picture of a large scale shock due to ram pressure located at the east of the galaxy possibly associated with particle acceleration. We argue that the asymmetry of the 6~cm polarized emission is characteristic for cluster galaxies and is a sign of interaction with its cluster environment. We conclude that it is probable that the galaxy experiences active ram pressure, but it is not clear whether we are observing it before, during, or after peak ICM pressure." }, "0403/astro-ph0403324_arXiv.txt": { "abstract": "I briefly review our current understanding of dark matter and dark energy. The first part of this paper focusses on issues pertaining to dark matter including observational evidence for its existence, current constraints and the `abundance of substructure' and `cuspy core' issues which arise in CDM. I also briefly describe MOND. The second part of this review focusses on dark energy. In this part I discuss the significance of the cosmological constant problem which leads to a predicted value of the cosmological constant which is almost $10^{123}$ times larger than the observed value $\\la/8\\pi G \\simeq 10^{-47}$GeV$^4$. Setting $\\la$ to this small value ensures that the acceleration of the universe is a fairly recent phenomenon giving rise to the `cosmic coincidence' conundrum according to which we live during a special epoch when the density in matter and $\\la$ are almost equal. Anthropic arguments are briefly discussed but more emphasis is placed upon dynamical dark energy models in which the equation of state is time dependent. These include Quintessence, Braneworld models, Chaplygin gas and Phantom energy. Model independent methods to determine the cosmic equation of state and the Statefinder diagnostic are also discussed. The Statefinder has the attractive property $\\atridot/a H^3 = 1 $ for LCDM, which is helpful for differentiating between LCDM and rival dark energy models. The review ends with a brief discussion of the fate of the universe in dark energy models. ", "introduction": "\\index{dark matter} Observations of the cosmic microwave background (CMB) and the deuterium abundance in the Universe suggest that $\\om_{\\rm baryon}h^2 \\simeq 0.02$, or $\\om_{\\rm baryon} \\simeq 0.04$ if the current Hubble expansion rate is $h = H_0/100 {\\rm km/sec/Mpc} = 0.7$. Although $\\om_{\\rm baryon}$ is much larger than the observed mass in stars, $\\om_{\\rm stars} \\simeq 0.005$ \\footnote{This suggests that most of the baryonic matter at $z=0$ is not contained in stars but might be contained in hot gas \\cite{bosma03}.}, it is nevertheless very much smaller than the total energy density in the universe inferred from the observed anisotropy in the cosmic microwave background \\cite{spergel03} \\beq \\Omega_{\\rm total} \\equiv \\frac{8\\pi G\\rho_{\\rm total}}{3H^2} = 1.02 \\pm 0.02~. \\eeq Both dark matter and dark energy are considered essential missing pieces in the cosmic jigsaw puzzle \\beq \\om_{\\rm total} - \\om_{\\rm baryons} = ~? \\eeq Although the nature of neither dark matter (DM) nor dark energy (DE) is currently known, it is felt that both DM and DE are non-baryonic in origin, and that DM is distinguished from DE by the fact that the former clusters on sub-Megaparsec scales (in order to explain galactic rotation curves) whereas the latter has a large negative pressure (and can make the universe accelerate). In addition there is strong evidence to suggest that \\beq \\Omega_m \\simeq 1/3, ~~ \\Omega_{\\rm DE} \\simeq 2/3~. \\eeq \\n In these lectures I will briefly review some properties of both dark matter and dark energy. Though the observational evidence favouring a flat Universe with $\\Omega_{\\rm total} \\simeq 1$ is fairly recent, the nature of the `unseen' component of the universe (which dominates its mass density), is a long-standing issue in modern cosmology. Indeed, the need for dark matter was originally pointed out by Zwicky (1933) who realized that the velocities of individual galaxies located within the Coma cluster were quite large, and that this cluster would be gravitationally bound only if its total mass substantially exceeded the sum of the masses of its component galaxies. For clusters which have relaxed to dynamical equilibrium the mean kinetic and potential energies are related by the virial theorem \\cite{coles} \\beq K + \\frac{U}{2} = 0~, \\eeq where $U \\simeq -GM^2/R$ is the potential energy of a cluster of radius $R$, $K \\simeq 3 M\\langle v_r^2\\rangle/2$ is the kinetic energy and $\\langle v_r^2\\rangle^{1/2}$ is the dispersion in the line-of-sight velocity of cluster galaxies. (Clusters in the Abell catalogue typically have $R \\simeq 1.5 h^{-1}$ Mpc.) This relation allows us to infer the mean gravitational potential energy if the kinetic energy is accurately known. The mass-to-light ratio in clusters can be as large as $M/L \\simeq 300 M_\\odot/L_\\odot$. However since most of the mass in clusters is in the form of hot, x-ray emitting intracluster gas, the extent of dark matter in these objects is estimated to be $M/M_{\\rm lum} \\simeq 20$, where $M_{\\rm lum}$ is the total mass in luminous matter including stars and gas. \\begin{figure}[ht] \\centering \\includegraphics[width=10cm]{five.eps} \\caption{\\footnotesize The observed rotation curve of the dwarf spiral galaxy M33 extends considerably beyond its optical image (shown superimposed); from Roy ~\\protect\\cite{dproy}. } \\label{fig:flat} \\end{figure} In individual galaxies the presence of dark matter has been convincingly established through the use of Kepler's third law \\beq v(r) = \\sqrt{\\frac{G M(r)}{r}} \\eeq to determine the `rotation curve' $v(r)$ at a given radial distance from the galactic center. Observations of galaxies taken at distances large enough for there to be no luminous galactic component indicate that, instead of declining at the expected rate $v \\propto r^{-1/2}$ true if $M \\simeq {\\rm constant}$, the velocity curves flattened out to $v \\simeq {\\rm constant}$ implying $M(r) \\propto r$ (see fig \\ref{fig:flat}). This observation suggests that the mass of galaxies continues to grow even when there is no luminous component to account for this increase. Velocity curves have been compiled for over 1000 spiral galaxies usually by measuring the 21 cm emission line from neutral hydrogen (HI) \\cite{dark1,dark2}. The results indicate that $M/L = (10 - 20) M_\\odot/L_\\odot$ in spiral galaxies and in ellipticals, while this ratio can increase to $M/L \\simeq (200 - 600) M_\\odot/L_\\odot$ in low surface brightness galaxies (LSB's) and in dwarfs. For instance, a recent measurement of the Draco dwarf spheroidal galaxy located at a distance of only 79 kpc from the Milky Way shows the presence of a considerable amount of dark matter $M/L\\vert_{\\rm Draco} = 440 \\pm 240 M_\\odot/L_\\odot$ \\cite{kleina01} ! It is interesting that the total mass of an individual galaxy is still somewhat of an unknown quantity since a turn around to the $v \\propto r^{-1/2}$ law at large radii has not been convincingly observed. An important difference between the distribution of dark matter in galaxies and clusters needs to be emphasised: whereas dark matter appears to {\\em increase} with distance in galaxies, in clusters exactly the reverse is true, the dark matter distribution actually {\\em decreases} with distance. Indeed, for certain dwarfs (such as DD0154) the rotation curve has been measured to almost 15 optical length scales indicating that the dark matter surrounding this object is extremely spread out (see also figure \\ref{fig:flat}). A foreground cluster, on the other hand, acts as a gravitational lens which focuses the light from background objects such as galaxies and QSO's thereby allowing us to determine the depth of the cluster potential well. Observations of strong lensing by clusters indicate that dark matter is strongly concentrated in central regions with a projected mass of $10^{13} - 10^{14} M_\\odot$ being contained within 0.2 - 0.3 Mpc of the central region. As we shall see later, this observation may prove to be problematic for alternatives to the dark matter hypothesis such as the Modified Newtonian Dynamics (MOND) approach of Milgrom \\cite{milgrom}. As discussed earlier, the fact that only $4\\%$ of the cosmic density is baryonic suggests that the dark matter which we are observing could well be non-baryonic in origin. The need for non-baryonic forms of dark matter gets indirect support from the fact that baryonic models find it difficult to grow structure from small initial conditions and hence to reconcile the existence of a well developed cosmic web of filaments, sheets and clusters at the present epoch with the exceedingly small amplitude of density perturbations ($\\delta\\rho/\\rho \\sim 10^{-5}$ at $z \\simeq 1,100$) inferred from COBE measurements and more recent CMB experiments. Indeed, it is well known that, if the effects of pressure are ignored, linearized density perturbations in a spatially flat matter dominated universe grow at the rate $\\delta \\propto t^{2/3} \\propto (1+z)^{-1}$, where $1+z = a_0/a(t)$ is the cosmological redshift. (Contrast this relatively slow growth rate with the exponential `Jeans instability' of a static matter distribution $\\delta \\propto \\exp{\\sqrt{4\\pi G\\rho} t}$.) In a baryonic universe, the large radiation pressure (caused by thompson scattering of CMB photons off electrons) ensures that density perturbations in the baryonic component can begin growing only after hydrogen recombines at $z \\simeq 1,100$ at which point of time baryons and radiation decouple. Requiring $\\delta > 1$ today implies $\\delta > 10^{-3}$ at recombination, which contradicts CMB observations by over an order of magnitude ! In non-baryonic models on the other hand, the absence of any significant coupling between dark matter and radiation allows structure to grow much earlier, significantly before hydrogen in the universe has recombined. After recombination baryons simply fall into the potential wells created for them by the dominant non-baryonic component. As a result a universe with a substantial non-baryonic component can give rise to the structure which we see today from smaller initial fluctuations. The growth of structure via gravitational instability depends both upon the nature of primordial perturbations (adiabatic/isocurvature) and upon whether the dark matter species is {\\rm hot} or {\\rm cold}. The issue of density perturbations has been discussed in considerable detail by Ruth Durrer at this school and I shall not touch upon this important topic any further. Let me instead say a few words about hot and cold dark matter. Non-baryonic {\\em Hot Dark Matter} (HDM) particles are assumed to have decoupled from the rest of matter/radiation when they were relativistic and so have a very large velocity dispersion (hence called `hot'). {\\em Cold Dark Matter} (CDM) particles, on the other hand, have a very small velocity dispersion and decouple from the rest of matter/radiation when they are non-relativistic. The free-streaming (collisionless phase mixing) of non-baryonic particles as they travel from high density to low density regions (and vice versa) introduces an important length scale called the `free-streaming distance' $\\lambda_{\\rm fs}$ -- which is the mean distance travelled by a relativistic particle species until its momentum becomes non-relativistic. In both HDM and CDM the processed {\\em final} spectrum of density perturbations differs from its initial form. In the case of HDM this difference arises because fluctuations on scales smaller than $\\lambda_{\\rm fs}$ are wiped out due to free streaming with the result that the processed final spectrum has a well defined cutoff on scales smaller than $\\lambda \\sim \\lambda_{\\rm fs}$. Perhaps the best example of HDM is provided by a light neutrino of mass about $30$ eV. In this case $\\lambda_{\\rm fs} \\simeq 41(30 {\\rm eV}/m_\\nu)$ Mpc with the result that large proto-pancakes having masses comparable to those of rich clusters of galaxies $M \\sim 10^{15} M_\\odot$ are the first objects to form in HDM. Smaller objects (galaxies) are formed by the fragmentation of the proto-pancake. This {\\em top-down} scenario for structure formation was originally suggested by Zeldovich and coworkers in connection with adiabatic baryonic models and subsequently applied to HDM. It has since fallen out of favour mainly due to the strong observational constraints on the mass of the neutrino $\\sum_{\\nu_i}m_{\\nu_i} < 0.7$ eV and on the relic neutrino density $10^{-3} \\lleq \\Omega_\\nu h^2 \\lleq 10^{-1}$ \\cite{elgaroy02,spergel03,ellis03,minaka02}. It also faces considerable difficulty in forming structure sufficiently early to explain the existence of galaxies and QSO's at high redshifts. In contrast to HDM, constituents of CDM have a much smaller free-streaming distance. Because of this small scales are the first to go non-linear and gravitational clustering proceeds in a {\\em bottom up} fashion in this scenario. A key quantity defining gravitational clustering is the power spectrum of density perturbations $P(k) \\equiv |\\delta_k|^2$, which is related to the mean square density fluctuation via \\beq \\bigg\\langle \\left (\\frac{\\delta\\rho}{\\rho}\\right )^2 \\bigg\\rangle = 4\\pi \\int_0^\\infty P(k) k^2 dk~. \\label{eq:pow1} \\eeq Inflationary models predict $P_i(k) \\propto k^n$, $n \\simeq 1$, at an early epoch. As the universe expands the power spectrum gets modified. The `processed' final spectrum depends upon the nature of dark matter, the epoch of matter-radiation equality and other cosmological quantities. The final and initial spectra are related through a transfer function \\beq P_f(k) = P_i(k)\\times T^2(k)~. \\label{eq:pow_spec} \\eeq CDM-type spectra have the following approximate form of the transfer function \\cite{sahni84,ss84,sc95} \\beq T(k) = \\left (1 + \\frac{A k^2}{\\log{(1 + Bk)}}\\right )^{-1}~. \\label{eq:transfer} \\eeq Equations (\\ref{eq:pow_spec}) \\& (\\ref{eq:transfer}) illustrate the `turn around' of the power spectrum from its primordial scale invariant form $P(k) \\propto k$ on the largest scales to $P(k) \\propto k^{-3}\\log^2k$ on small scales. (The precise location of the turn-around and the amplitude of $P(k)$ depend upon specific details of the cosmological model, see for instance \\cite{bbks86}.) The `standard' cold dark matter (SCDM) paradigm, which assumed that $\\Omega_{\\rm CDM} = 1$, was introduced during the early 1980's at roughly the same time when HDM was perceived to be in trouble (see \\cite{kolbt90,khlopov,coles,sc95} for references to earlier work on this subject). Although SCDM was very successful in explaining a host of observational details, it was clear already a decade ago, that the processed power spectrum of SCDM lacked sufficient power on large scales and so fell short of explaining the angular two point correlation function for galaxies on scales $\\sim 50$ Mpc \\cite{efstath}. The relevant cosmological quantity in this case is the shape of the power spectrum of density perturbations, which for CDM-like models, can be characterised by the `shape parameter' $\\Gamma = \\Omega_m h$. SCDM models with $\\Omega_m = 1$ and the HST-determined value $h \\simeq 0.7$ predict $\\Gamma \\simeq 0.5$ which is much larger than the observed value $\\Gamma = 0.207 \\pm 0.030$ inferred from observations of galaxy clustering in the sloan digital sky survey (SDSS) \\cite{pope}. A modification of SCDM called LCDM assumes that, in addition to CDM the universe consists of a smoothly distributed component called a cosmological constant or a Lambda-term. LCDM models with $h \\simeq 0.7$ and $\\Omega_m = 0.3$ predict a smaller value for the shape parameter, $\\Gamma \\simeq 0.2$, and the resulting amplitude and shape of the power spectrum is in excellent agreement with several different sets of observations as demonstrated in figure \\ref{fig:power}. \\begin{figure}[ht] \\centering \\includegraphics[width=10cm]{kspace.ps} \\caption{\\footnotesize The power spectrum inferred from observations of large scale structure, the Lyman$\\alpha$ forest, gravitational lensing and the CMB. The solid line shows the power spectrum prediction for a flat scale-invariant LCDM model with $\\Omega_m = 0.28$, $\\Omega_b/\\Omega_m = 0.16$, $h = 0.72$; from Tegmark et al. ~\\protect\\cite{tegmark03a}. } \\label{fig:power} \\end{figure} From (\\ref{eq:pow1}), (\\ref{eq:pow_spec}) \\& (\\ref{eq:transfer}) we find that on small scales, the contribution to the {\\em rms} density fluctuation from a given logarithmic interval in $k$ is \\beq \\left (\\frac{\\delta\\rho}{\\rho}\\right )^2_k \\sim k^3 P_f(k) \\propto \\log^2{k}~, \\eeq which illustrates the fact that, although the smallest scales are the first to go non-linear, there is significant power to drive gravitational instability rapidly to larger scales in this model. Indeed, detailed N-body simulations of large scale structure show that filaments defining the cosmic web first form on the smallest scales. The scale-length characterizing the cosmic web grows as the universe expands, until at the present epoch the cosmic web consists of a fully developed supercluster-void network with a scale-length of several tens of Megaparsec \\cite{sss96,sss03,masaar,rien}. Promising candidates for cold dark matter include a $100 - 1000$ GeV particle called a neutralino. The neutralino is a weakly interacting massive particle (WIMP). As its name suggests it is neutral and is a fermionic partner to the gauge and Higgs bosons (usually called the `bino, wino and higgsino'). It is believed that the lightest supersymmetric particle will be stable due to R-parity which makes the neutralino an excellent candidate for cold dark matter (see \\cite{roszko99,jungman96} for reviews of particle dark matter). A radically different particle candidate for cold dark matter is an ultra-light pseudo-Goldstone boson called an axion with a mass of only $m_a \\sim 10^{-5 \\pm 1}$ eV. Although ultralight, the axion is `cold' because it was created as a zero-momentum condensate. Its existence is a by-product of an attempt to resolve QCD of what is commonly called the `strong CP problem' which arises because non-perturbative effects in QCD give rise to an electric dipole moment for the neutron -- in marked contrast with observations \\cite{kolbt90}. Other candidates for non-baryonic cold dark matter include string theory motivated modulii fields \\cite{brustein}; non-thermally produced super-heavy particles having a mass $\\sim 10^{14}$ GeV and dubbed Wimpzillas \\cite{kolb}; as well as axino's and gravitino's -- superpartners of the axion and graviton respectively \\cite{roszko99}. Since WIMP's cluster gravitationally, one should expect to find a flux of these particles in our own solar system and attempts are being made to determine dark matter particles by measuring the scattering of WIMP's on target nucleii through nuclear recoils. Now the earth orbits the sun with a velocity $\\simeq 30$ km/sec, even as the sun orbits the galaxy with $v_{M_\\odot} \\simeq 220$ km/sec. Furthermore the plane of the Earth's orbit is inclined at an angle of $60^\\circ$ to the glactic plane, because of which the dark matter flux on Earth is expected to be larger in June (when the Earth's velocity and the Sun's velocity add together) than in December (when these two velocities subtract). The resulting rate variation is about $7\\%$ between the flux measured during summer and winter. Precisely such a signal was reported by the DAMA experiment whose data (collected since 1996) appears to show a yearly modulation with greater events reported in June than in December \\cite{dama}. However results obtained by the DAMA group remain controversial since they have not been substantiated by other groups which report negative results for similar searches (see \\cite{munoz,khalil} for recent reviews on this subject). \\begin{figure}[ht] \\centering \\includegraphics[width=5cm]{sol.eps} \\caption{\\footnotesize The Earth's motion around the Sun; from Khalil and Munoz (2001). } \\bigskip \\medskip \\label{fig:sol} \\end{figure} Despite the excellent agreement of LCDM with large scale observations, some concerns have recently been expressed about the ability of this model to account for a number of smaller scale observations which can be summarized as follows: \\begin{itemize} \\item {\\em The substructure problem} is used to describe the fact that the cold dark matter model (with or without a cosmological constant) predicts an excessive number of dark matter subhaloes (or substructure) within a larger halo. If one (perhaps naively) associates each halo with a gravitationally bound baryonic object then the predicted number of dwarf-galaxy satellites within the local group exceeds the observed number by over an order of magnitude. Indeed, detailed N-body simulations as well as theoretical estimates predict around 1000 dark matter satellites in our local group which is much larger than the 40 or so observed at present \\cite{klypin99,moore99,kauffmann93,bullock00, somerville02,benson02,mateo98,tasi02,evans01}. \\item {\\em The cuspy core problem} CDM predicts a {\\em universal density profile} for dark matter halos in the wide range $10^7 M_\\odot - 10^{15} M_\\odot$ which applies both to galaxy clusters as well as individual galaxies including dwarfs and LSB's. \\footnote{Low Surface Brightness Galaxies (LSB's) are dominated by their dark matter content and therefore provide particularly good astrophysical objects with which to test dark matter models.} The density profile originally suggested by Navarro, Frenk and White \\cite{nfw} is \\beq \\rho(r) = \\rho_0\\left (r_s/r\\right )\\bigg\\lbrack 1 + \\left (\\frac{r}{r_s}\\right )\\bigg \\rbrack^{-2}~, \\eeq which gives $\\rho \\propto r^{-1}$ for $r \\ll r_s$ and $\\rho \\propto r^{-3}$ for $r \\gg r_s$, where $r_s$ is the scale radius and $\\rho_0$ is the characteristic halo density. (Other groups using higher resolution computations found somewhat steeper density profiles at small radii, such as $\\rho \\propto r^{-1.5}$ \\cite{moore99b,jing00}.) {\\em The cuspy core problem} refers to the apparent contradiction between N-body experiments -- which show that the density profile in CDM halos has a $1/r$ (or steeper) density cusp at the center, and observations -- which appear to favour significantly shallower density cores in galaxy clusters as well as in individual dwarf and LSB galaxies (see \\cite{flores94,burkert95,bolatto02,blok97,primack01,tasi02,klypin04,rhee03,sand03,ma04} for detailed discussions of this issue). \\end{itemize} Although disconcerting, given the very considerable success of LCDM in explaining gravitational clustering on large scales, it may at this point be premature to condemn this model on the basis of small scale observations alone. It could be that the difficulties alluded to above are a result of an oversimplification of the complex physical processes involved and that a more careful analysis of the baryonic physics on small scales including the hydrodynamical effects of star formation and supernova feedback needs to be undertaken. For instance both dwarfs and LSB's have very shallow potential wells, a strong burst of star formation and supernova activity may therefore empty dark matter halos of their baryonic content resulting in a large number of `failed galaxies' and providing a possible resolution to the `satellite catastrophe'. (The failed galaxies will act as gravitational lenses and should therefore be detectable through careful observations.) Other explanations include the effects of tidal stripping recently discussed in \\cite{klypin04}. Likewise issues involving beam smearing, the influence of bars and the interaction of baryons and dark matter in the central regions of galaxies and clusters could be intricately linked with the central cusp issue and must be better understood if one wishes to seriously test the CDM hypothesis on small scales. In concluding this discussion on dark matter I would like to briefly mention Modified Newtonian Dynamics (MOND) which, in some circles, is regarded as an alternative to the dark matter hypothesis. As the name suggests, MOND is a modification of Newtonian physics which proposes to explain the flat rotation curves of galaxies without invoking any assumptions about dark matter. Briefly, MOND assumes that Newtons law of inertia ($F = ma$) is modified at sufficiently low accelerations ($a < a_0$) to \\beq {\\bf F} = m{\\bf a} \\mu(a/a_0)~, \\eeq where $\\mu(x) = x$ when $x \\ll 1$ and $\\mu(x) = 1$ when $x \\gg 1$ \\cite{milgrom,sanders02}. It is easy to see that this results in the modification of the conventional formula for gravitational acceleration ${\\bf F} = m{\\bf g_N}$, resulting in the following relation between the true acceleration and the Newtonian value: $a = \\sqrt{g_N a_0}$. For a body orbiting a point mass $M$, $g_N = GM/r^2$. Since the centripetal acceleration $a = v^2/r$ now equals the {\\em true} acceleration $a$, one gets \\beq v^4 = GMa_0~, \\eeq \\ie for sufficiently low values of the acceleration the rotation curve of an isolated body of mass $M$ does not depend upon the radial distance $r$ at which the velocity is measured, in other words not only does this theory predict flat rotation curves it also suggests that the individual halo associated with a galaxy is infinite in extent ! (This latter prediction may be a problem for MOND since recent galaxy-galaxy lensing results \\cite{hoekstra02} suggest that galaxy halo's may have a maximum extent of about 0.5 Mpc.) The value of $a_0$ needed to explain observations is $a_0 \\sim 10^{-8}$cm/s$^2$ which is of the same order as $cH_0$ ! This has led supporters of this hypothesis to conjecture that MOND may reflect ``the effect of cosmology on local particle dynamics'' \\cite{sanders02}. Although MOND gives results which are in good agreement with observations of individual galaxies, it is not clear whether it is as successful for explaining clusters for which strong gravitational lensing indicates a larger mass concentration at cluster centers than accounted for by MOND \\cite{sanders02,combes02}. Another difficulty with MOND is that it is problematic to embed this theory within a more comprehensive relativistic theory of gravity and hence, at present, it is not clear what predictions a MOND-type theory may make for gravitational lensing and other curved space-time effects. For some recent developments in this direction see \\cite{beken04}. To summarise, current observations make a strong case for clustered, non-baryonic dark matter to account for as much as a third of the total matter density in the Universe $\\Omega_m \\simeq 1/3$. The remaining two-thirds is thought to reside in a relative smooth component having large negative pressure and called Dark Energy. ", "conclusions": "From the theoretical standpoint the single most important question to be asked of dark energy is \\centerline{Is $w = -1$ ?} \\smallskip \\n Rephrased in terms of the Statefinder diagnostic the question is: \\smallskip \\centerline{Is $\\atridot/a H^3 = 1$ ?} \\smallskip \\n If future observations do answer this question in the affirmative\\footnote{\\ie if $w = -1$ is measured to satisfyingly high accuracy} then, in all likelyhood the cosmological constant is the vacuum energy, and one will need to review the cosmological constant problem again, in order to fathom why the formally infinite quantity $\\tik$ is in fact so very small. \\begin{figure}[ht] \\centering \\includegraphics[width=9cm]{confcmbclust.eps} \\caption{\\footnotesize Target statistical uncertainty of the SNAP experiment is shown overlayed with current results from CMB and LSS observations. From Aldering \\cite{aldering}. } \\bigskip \\medskip \\label{fig:snap} \\end{figure} If on the other hand, either $w \\neq -1$ or if the DE density is shown to be time dependent, then the cosmological constant problem may need to be decoupled from the DE conundrum and searches for evolving DE models which produce $\\rho_{\\rm DE} \\simeq 10^{-47}$GeV$^4$ without exacerbating `cosmic coincidence' will need to be examined deeply in the light of developments both in high energy physics and in gravitation theory (superstring/M-theory, extra dimensions etc.). In either case the key to determining the properties of DE to great precision clearly lies with ongoing and future astrophysical experiments and observations. Since the original discovery of an accelerating universe \\cite{perl98a,perl98b,riess98} the Sn data base has grown considerably and data pertaining to $\\sim 200$ type Ia supernovae are avaliable in the literature \\cite{tonry03,knop03,barris03,riess04}. Although systematic effects such as luminosity evolution, dimming by intervening extragalactic material (alternatively brightening due to gravitational lensing) continue to be a cause of some concern -- recall that a luminosity evolution of $\\sim 25\\%$ over a lookback time of $\\sim 5$ Gyr is sufficient to nullify the DE hypothesis \\cite{riess99} -- it is reassuring that recent observations of CMB anisotropies and estimates of galaxy clustering in the 2dF and SDSS surveys, make a strong and independent case for dark energy \\cite{spergel03,tegmark03a,tegmark03b}. Indeed, a joint analysis of CMB data from WMAP + HST Key Project determination of $H_0$ imply $w_{\\rm DE} < -0.5$ at the $95\\%$ confidence level \\cite{spergel03}. It is of paramount importance that Sn observations continue to be supplemented by other investigations which are sensitive to the geometry of space and can be used as independent tests of the DE hypothesis. The volume-redshift test, Sunyaev-Zeldovich surveys, the Alcock-Paczynski test, the angular size-redshift test and gravitational lensing have all been suggested as possible probes of dark energy, and will doubtless enrich the theory vs observations debate in the near future. In addition, the proposed SNAP satellite which aims to measure light curves of $\\sim 2000$ supernovae within a single year \\cite{snap}, should provide a big step forward in our understanding of type Ia supernovae and help determine the cosmological parameters to great precision, as shown in figure \\ref{fig:snap}." }, "0403/astro-ph0403438_arXiv.txt": { "abstract": "Of all pulsars known Vela has been one of the most productive in terms in understanding pulsars and their characteristics. We present the latest results derived from Australian telescopes. These include a more accurate pulsar distance, a more precise pulsar local space velocity, a new model of the spin up and the association of a radio nebula with the X-ray pulsar wind nebula. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403112_arXiv.txt": { "abstract": "We present \\hst\\ observations that show a bifurcation of colors in the middle main sequence of the globular cluster $\\omega$ Centauri. We see this in three different fields, observed with different cameras and filters. We also present high precision photometry of a central ACS field which shows a number of main-sequence turnoffs and subgiant branches. The double main sequence, the multiple turnoffs and subgiant branches, and other population sequences discovered in the past along the red giant branch of this cluster add up to a fascinating but frustrating puzzle. We suggest various explanations, none of them very conclusive. ", "introduction": "A number of properties (total mass, chemical composition, kinematics, and spatial distribution of the stars) make $\\omega$ Centauri a peculiar object among Galactic globular clusters. The most evident anomaly is the large spread in metallicity seen both in spectroscopic (Norris \\& Da Costa 1995) and photometric (Hilker \\& Richtler 2000, Lee et al.\\ 1999, Pancino et al.\\ 2000) investigations. Most of the fascinating results on \\omcen~ come from the evolved stellar population, which can be studied in detail from the ground. In this paper we use \\hst~ data to explore the cluster's turnoff (TO) and main-sequence (MS) populations. While studies of the evolved red giant branch (RGB) can explore metallicity, kinematic, and spatial-distribution issues, we need the fainter stars if we hope to learn anything about ages and mass functions, and to give us better statistics (there are $\\sim$10 MS stars for every RGB star). The present paper was stimulated by preliminary results by one of us (Anderson 1997, 2002, 2003), on the presence of multiple turnoffs and of a bifurcated main sequence (MS). Here we confirm that the unusual features found in the color-magnitude diagrams are not some data-reduction artifact, or a local phenomenon. The features are real and are present throughout the cluster. Unfortunately, the striking results we present here lead to more questions than they answer. Though it will take a lot of time to fully exploit the \\omcen~ data stored in the \\hst~ archive (and we are working on this), we think that these new results are worthy of immediate publication because of their importance to the ongoing debate on the nature of this object. ", "conclusions": "One way to interpret the observations is to assume that either the photometric calibration or the isochrones are in error. If the net error is 0.06 mag in $V-I$ color, then perhaps the metal-poor (MP) population ([Fe/H]\\ $=\\!-1.6$) follows along the bMS instead of along the rMS. If this is the case then the rMS would correspond to the metal-rich population ([Fe/H]\\ $=\\!-0.5$), as the 0.06 mag $V\\!-\\!I$ separation cannot be explained by the metallicity difference between the MP and Mint populations. (While it may be conceivable that the isochrones could have errors in an absolute sense, they should be reliable in a differential sense.) There are additional problems with this interpretation. First is that only 5$\\%$ of the RGB stars are metal rich, but in this scenario over 70$\\%$ of the MS stars would be metal rich. This would imply drastically different mass functions, such as have never been seen before anywhere (see Piotto \\& Zoccali 1999). Furthermore, there is no actual gap in the observed metallicity distribution and in the color distribution of the RGBs of the MP and Mint populations. Most importantly, the fact that the MS extension of the LTO runs parallel to the rMS (on the red side of it, panel e) makes this scenario impossible. The second interpretation is that the rMS corresponds to the MP stars, but the bMS corresponds to a super-metal-poor population, with [Fe/H] $\\ll\\!-2$. However, such a large population of metal-poor stars has never been observed in \\omcen\\ or in any other globular. The third possibility is that the populations of the two MSs have sensibly different helium content ($Y$). Norris, Freeman, \\& Mighell (1996) have shown that the metallicity distribution of \\omcen\\ stars can be well fitted by two separate components, and argued that this can be explained by two successive epochs of star formation. Assuming for the more metal-rich ([Fe/H]\\ $=\\!-1.0$) Mint population a helium content of $Y\\sim0.30$, we find that the corresponding MS would be $\\sim0.07$ magnitude bluer in ($V-I$) than the MP MS (assumed to have a canonical $Y=0.23$, and [Fe/H]\\ $=\\!-1.6$). Note that Norris et al.\\ (1996) found that the ratio of the Mint to MP population should be 0.2, compatible, within the uncertainties, with the value we find for the rMS/bMS ratio. Panel e of Fig.\\ 1 shows that the bMS could well be connected with the intermediate TO-SGB. Panel a of Fig.\\ 2 shows that this intermediate SGB is slightly brighter than the luminosity expected for a metallicity similar to the Mint, and the expected TO is redder than the observed one. These observational facts are consistent with this population beeing helium enhanced and slightly younger, as expected if the helium enhancement is due to self-pollution from intermediate AGB MP stars. The dramatic increase of s-process heavy-element abundances with metallicity found by Smith et al.\\ (2000) in \\omcen\\ RGB stars furtherly support the hypothesis that Mint stars could have formed from material polluted by ejecta from 1.5-3 $m_\\odot$ AGB stars. The presence of a population with high helium content could also account for the anomalously hot HB of \\omcen, following the calculations of D'Antona et al.\\ (2002). All this notwithstanding, a $Y\\geq0.30$ is higher than any value so far measured in Galactic GCs (Salaris et al.\\ 2004), and not easy to understand. As a fourth possibility, if we assume that the rMS corresponds to the majority of the cluster stars, the bMS could correspond to a population of stars located behind \\omcen. As shown in panel d of Fig.\\ 2, if the bMS is populated by stars located 1.6 kpc beyond \\omcen, we can easily fit it with an [Fe/H] $=\\!-1$ isochrone. Panel e of Fig.\\ 1 appears to strengthen this hypothesis: we see the bMS get closer and closer to the rMS, crossing it at $H\\alpha\\sim18.5$, and apparently continuing into a broadened TO and SGB. This broadening of the intermediate TO could be the result of a spread in both metallicity and distance. The overall appearance of the CMD is that there are two sequences, shifted by up to $\\sim0.3$--0.5 magnitude. The hypothesis of a background agglomerate of stars with metallicity around [Fe/H] $\\sim\\!-1.0$ would also naturally explain why the bMS appears to intersect the rMS at $V_{606}\\sim23.5$ (cf.\\ Figs.\\ 1c and 1d). Such a background object would naturally explain the observation that the giants of different metallicity appear to have somewhat different spatial distributions (Jurcsik 1998, Hilker \\& Richtler 2000), though this spatial variation could be explained by merger or self-enrichment scenarios as well. Leon, Meylan, \\& Combes (2000) have identified a tidal tail around \\omcen. Tidal tails often have a clumpy nature. However, the number of stars in the bMS seems to be too large and the sequence too sharp to be interpreted as a part of a clump in a tidal tail behind the cluster. Another possibility is that the object in the background is a distinct cluster or a dwarf galaxy. As it should cover at least 20--30 arcmin in the sky (this is the extent of the region where we identified a DMS) and be located at about 7 kpc from the Sun, the object should be extended by at least 40--60 pc. The probability of observing such an object in the direction of \\omcen~ is extremely low. However, if this object happens to be gravitationally linked to \\omcen~(either because it was part of the same original system or because it is the remnant of some merging event), that would enhance the probability of seeing it in the same direction as \\omcen. We note that the idea of a population of stars behind the cluster has been suggested before. Ferraro et al.\\ (2002) measured a bulk motion for the RGB-a stars with respect to the other cluster stars, and interpreted this as evidence that it could be a background object, or a merger product that has not yet phase-mixed. However, Platais et al.\\ (2003) find this motion spurious, attributing it to a color/magnitude term in the proper motions. Moreover, Anderson (2003), using very accurate WFPC2 proper motions, contradicts the bulk motions seen by Ferraro et al.\\ (see his Fig.\\ 1). In any case, the background population we consider here could not correspond to the very metal-rich population; our Fig.\\ 1e makes it clear that the LTO and the bMS are not related to each other." }, "0403/astro-ph0403397_arXiv.txt": { "abstract": "Using a sample of 2408 time--resolved spectra for 91 BATSE GRBs presented by Preece et al., we show that the relation between the isotropic--equivalent luminosity ($L_{\\rm{iso}}$) and the peak energy ($E^{'}_{\\rm{p}}$) of the $\\nu F_{\\nu}$ spectrum in the cosmological rest frame, $L_{\\rm{iso}}\\propto E_{\\rm{p}}^{'2}$, holds within these bursts, and also holds among these GRBs, assuming that the burst rate as a function of redshift is proportional to the star formation rate. The possible implications of this relation for the fireball models are discussed by defining a parameter $\\omega\\equiv (L_{\\rm{iso}}/10^{52} {\\rm{erg\\, s}^{-1}})^{0.5}/(E^{'}_{\\rm{p}}/200\\, {\\rm{ keV}})$. It is found that $\\omega$ is narrowly clustered in $0.1-1$. We constrain some parameters for both the internal shock and external shock models from the requirement of $\\omega\\sim 0.1-1$, assuming that these model parameters are uncorrelated. The distributions of the parameters suggest that if the prompt gamma--rays are produced from kinetic--energy--dominated internal shocks, they may be radiated from a region around $R\\sim 10^{12}-10^{13}$ cm (or Lorentz factor $\\sim 130-410$) with a combined internal shock parameter $\\zeta_{\\rm{i}}\\sim 0.1-1$ during the prompt gamma--ray phase, which are consistent with the standard internal shock model; if the prompt gamma--rays of these GRBs are radiated from magnetic--dissipation--dominated external shocks, the narrow cluster of $\\omega$ requires $\\sigma\\sim 1-470$, $\\Gamma\\sim 216-511$, $E\\sim10^{51}-10^{54}$ ergs, $n\\sim 0.5-470$ cm$^{-3}$, and $\\zeta_{\\rm{e}}\\sim 0.36-3.6$, where $\\sigma$ is the ratio of the cold-to-hot luminosity components, $\\Gamma$ the bulk Lorentz factor of the fireball, $E$ the total energy release in gamma--ray band, $n$ the medium number density, and $\\zeta_{\\rm{e}}$ a combined external shock parameter, which are also in a good agreement with the fittings to the afterglow data. These results indicate that both the kinetic--energy--dominated internal shock model and the magnetic--dissipation--dominated external shock model can well interpret the $L_{\\rm{iso}}\\propto E_{\\rm{p}}^{'2}$ relation and the value of $\\omega$. ", "introduction": "% Gamma--ray bursts (GRBs) are now believed to be produced by jets powered by central engines with a standard energy reservoir at cosmological distances (see series reviews by Fishman \\& Meegan 1995; Piran 1999; van Paradijs et al. 2000; Cheng \\& Lu 2001; M\\'{e}sz\\'{a}ros 2002; Zhang \\& M\\'{e}sz\\'{a}ros 2003). The most impressive features of GRBs are the great diversities of their light curves and spectral behaviors, and extremely large luminosities. These spectra are well fitted by the Band function (Band et al. 1993). However, the radiation mechanism at work during the prompt phase remains poorly understood. Although the spectral behavior and the luminosity are dramatically different from burst to burst, the isotropic--equivalent luminosity, $L_{\\rm{iso}}$, (or isotropic--equivalent energy radiated by the source, $E_{\\rm{iso}}$), and $E^{'}_{\\rm{p}}$, the peak energies of $\\nu F_{\\nu}$ spectrum in the rest frame among GRBs, obey an empirical relation of $L_{\\rm{iso}}\\propto E_{\\rm{p}}^{'2}$ (Amati et al. 2002; Yonetoku et al. 2003; Sakamoto et al. 2004; Lamb et al. 2003a, b, c). This relation was revisited in standard synchrotron/inverse--Compton/synchro--Compton models (Zhang \\& M\\'{e}sz\\'{a}ros 2002). Recently, Sakamoto et al. (2004) and Lamb et al. (2003a, b, c) pointed out that HETE--2 observations not only confirm this correlation, but also extend it to the population of X--ray flashes, which are thought to be a low energy extension of typical GRBs (Heise et al. 2001, Kippen et al. 2003). Based on this relation, Atteia (2003) also constructed a simple redshift indicator for GRBs. One may ask: whether or not this relation holds in any segment within a GRB? The answer remains unknown. If the answer is positive, combining the results mentioned above, one might suggest that this relation is a universal law during the prompt gamma--ray phase, and presents some constraints on fireball models. In this Letter, we investigate this issue. Using a sample of 2408 time--resolved spectra for 91 BATSE GRBs presented by Preece et al. (2000), we show that this relation holds within these bursts, and also holds among these GRBs, assuming that the burst rate as a function of redshift is proportional to the star formation rate. We suggest that both the kinetic--energy--dominated internal shock model and the magnetic--dissipation--dominated external shock model may well interpret this relation. Throughout this work we adopt $H_0=65$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{\\rm{m}}=0.3$, and $\\Omega_\\Lambda=0.7$. ", "conclusions": "Using a sample of 2408 time--resolved spectra for 91 long, bright GRBs presented by Preece et al. (2000), we show that the $L_{\\rm{iso}}\\sim E_{\\rm{p}}^{'2}$ relation holds within these BATSE bursts, and this relation also holds among these GRBs by assuming that the burst rate as a function of redshift is proportional to the star formation rate. We discuss possible implications of this relationship for the fireball models by defining a parameter $\\omega\\equiv (L_{\\rm{iso}}/10^{52} {\\rm{erg\\ s}^{-1}})^{0.5}/(E^{'}_{\\rm{p}}/200\\, {\\rm{keV}})$. It is found that $\\omega$ is not influenced by the Doppler--boosting effect, and it is determined by the gamma--ray emission region and shock parameters in the kinetic--energy--dominated internal shock model or determined by the parameters of both the shock and the environment in the magnetic--dissipation--dominated external shock model. We derive the distributions of some parameters for both the internal shock model and the external shock model from the requirement of $\\omega\\sim 0.1-1$. We suggest that if the prompt gamma--rays are produced from a kinetic--energy--dominated internal shock, they may be radiated from a region around $R\\sim 10^{12}-10^{13}$ cm (or Lorentz factor $\\sim 130-410$) with an internal shock parameter $\\zeta_{\\rm{i}}\\sim 0.1-1$, which is consistent with the standard internal shock model; if the prompt gamma--rays of these GRBs are radiated from magnetic--dissipation--dominated external shocks, the $\\omega\\sim 0.1-1$ requires $\\sigma \\sim 1-470$, $\\Gamma \\sim 216-511$, $\\zeta_{\\rm{e}}\\sim 0.36-3.6$, $E\\sim 10^{51}-10^{54}$ ergs, and $n\\sim 0.5-470$ cm$^{-3}$. Please note that the distributions for these model parameters for both the internal and external shock models are based on the assumption that they are uncorrelated. Although these parameters seem to be uncorrelated, we do not know if it is really the case. If these parameters are correlated during prompt gamma--ray phase, these distributions are not valid. We would like to thank the referee, Dr. Don Lamb, for his valuable comments, which have enabled us to improve greatly the manuscript. This work is supported by the National Natural Science Foundation of China (grants 10233010 and 10221001), the National 973 Project (NKBRSF G19990754), the Natural Science Foundation of Yunnan (2001A0025Q), and the Research Foundation of Guangxi University." }, "0403/astro-ph0403442_arXiv.txt": { "abstract": "The variability of the Fe \\Ka line near 6.5 keV seems to be reduced compared to the variability of the hard X-rays which presumably drive the line emission. This is observed both in active galactic nuclei and galactic black hole binaries. We point out that such reduced variability, as well as lack of coherence between the variations of the line and the continuum, are a natural prediction of a propagation model of variability in the geometry of inner hot accretion flow. We compute detail model predictions of the variability characteristics which could be compared with current and future data. We also point out that the model requires a gradual disappearance of the cold disc, rather than a sharp transition from the cold disc to a hot flow. ", "introduction": "The Fe \\Ka fluorescent/recombination line near 6.5 keV is an important diagnostic of accretion flows around compact objects (see Reynolds \\& Nowak 2003 for a recent review). It is the strongest hard X-ray ($E>1$ keV) line which can originate in the innermost regions of accretion flows ($\\le 100\\,\\Rg$). It is indeed observed in energy spectra from all kinds of accreting sources: black hole and neutron star X-ray binaries, cataclysmic variables, active galactic nuclei (AGN). The line is produced when plasma is irradiated by hard X-ray ($E > 7$ keV) radiation. If the plasma is Thomson thick the continuum spectral component formed as a result of the irradiation (``Compton reflection'') has a characteristic shape, peaking at 20--30 keV (Lightman \\& White 1988). The line and the Fe K-shell absorption edge are superposed on this continuum (George \\& Fabian 1991; Matt, Perola \\& Piro 1991). The properties of the line and edge depend on the ionization of the reflecting medium: with increasing ionization the line and edge shift towards higher energies, while their strength increase (e.g.\\ \\.{Z}ycki \\& Czerny 1994 and references therein). The profile of the line may be modified by relativistic and kinematic effects, if the line originates in e.g.\\ a rotating accretion disc (Fabian et al.\\ 1989). The line and absorption edge are then broadened and smeared. Such broad features are commonly seen in Seyfert galaxies (e.g.\\ MCG-6-30-15, Tanaka et al.\\ 1995, Fabian et al.\\ 2002) and black hole binaries (BHB; e.g.\\ Cyg X-1, Done \\& \\.{Z}ycki 1999; GRS 1915+105, Martocchia et al.\\ 2002), providing a clear evidence for a relativistic accretion disk extending deep into the gravitational potential of the central black hole. X-ray emission from accreting sources is highly variable in a broad range of time-scales. Most of the variability power is located in the range of Fourier frequency $f \\approx$(0.1--1)$M/(10\\,\\MSun)$ Hz, corresponding to a time-scale ($T=1/f$) of a few seconds for a $10\\,\\MSun$ stellar black hole and a few weeks for a $10^7\\,\\MSun$ AGN (BHB review in McClintock \\& Remillard 2003; Markowitz et al.\\ 2003b for AGN). Typical power density spectrum is roughly a power law with slope $\\alpha\\approx 0$ ($P(f)\\propto f^{\\alpha}$) at low $f$, steepening to $\\alpha\\approx -1$ at $f \\approx 0.1$ Hz and to $\\alpha\\approx -2$ at $f=1$--3 Hz, for the BHB in low/hard state. Typical root-mean-square (r.m.s.) variability is 20--30\\%. The variability is stochastic rather than caused by deterministic chaos type of process (Czerny \\& Lehto 1997). The observed variability of the Fe \\Ka line and the entire reprocessed component is somewhat surprising: they generally show rather less variability than the high energy continuum which is presumably driving the line emission and Compton reflection. This is seen both in AGN and BHB. In AGN the \\Ka line seems either not to respond to continuum variations on time-scales of minutes to days (e.g.\\ Reynolds 2000; Done, Madejski \\& \\.{Z}ycki 2000; Chiang et al.\\ 2000), or the line variability appears to be uncorrelated with that of the continuum (Vaughan \\& Edelson 2001). In particular, studies of r.m.s.\\ variability amplitude as a function of energy demonstrate the reduced variability in a relatively model independent way (Inoue \\& Matsumoto 2001; Markowitz, Edelson \\& Vaughan 2003a). The short term variability of the reprocessed component is more difficult to measure because of insufficient statistics, but where such studies were possible, the results also suggested reduced variability (e.g.\\ Done et al.\\ 2000). A similar effect is seen in BHB systems. The time resolved spectral analysis is more difficult for BHB than for AGN, but Fourier spectroscopy (study of Fourier-frequency resolved spectra; Revnivtsev, Gilfanov \\& Churazov 1999, 2001) can be used to investigate the variability characteristics on short time scales. Indeed, the amplitude of the reprocessed component in high Fourier frequency spectra is smaller compared to that in low Fourier frequency spectra (Revnivtsev et al.\\ 1999, 2001). In other words, the reprocessed component responds to the variability of the primary continuum on long time scales ($T=1/f \\ge 1$ sec) but it does not do so on short time scales. In AGN one reason for the lack of variability might be the contribution to the reprocessing by a distant matter, e.g.\\ the obscuring dusty torus. Indeed, a narrow Fe \\Ka line is observed in many Seyfert galaxies. However, it is the variability properties of the broad component (hence produced very close to the central black hole) that are so puzzling. One suggested explanation invoked complex ionization effects in the illuminated surface of the disc (e.g.\\ Nayakshin, Kazanas \\& Kallman 2000). Formation of a hot ionized skin (where the Fe \\Ka line is not produced) with thickness proportional to the X--ray flux may lead to anti-correlation between the line flux and the continuum flux. This model was quantitatively tested by \\.{Z}ycki \\& R\\'{o}\\.{z}a\\'{n}ska (2001), who concluded that it does indeed predict certain decrease of amplitude of variability of the line, although it is not possible to obtain an absolutely constant line flux. In this paper we demonstrate that the reduced variability of the Fe \\Ka line is a natural result from a propagation model of X-ray emission in the geometry of a hot inner accretion flow. This geometry is one of the possibilities for accretion flow in low/hard states of accreting black holes (e.g.\\ Di Salvo et al.\\ 2001; see Done 2002 for review). It is supported by X--ray spectral studies (review in Poutanen 1999), in particular the correlation between spectral slope and amplitude of reflection (Zdziarski, Lubi\\'{n}ski \\& Smith 1999) and long time scale spectral evolution of black hole binaries (Zdziarski et al.\\ 2003). The physical mechanism of formation the two-phase plasma flow may be related to plasma evaporation/condensation (R\\'{o}\\.{z}a\\'{n}ska \\& Czerny 2000 and references therein). The idea of propagating X-ray emitting structures was put forward by Miyamoto et al.\\ (1988), based on early variability studies of Cyg X-1. This was further developed and tested by e.g.\\ Nowak et al.\\ (1999), Misra (2000), and formulated in a more general form by Kotov, Churazov \\& Gilfanov (2001). Recently, \\.{Z}ycki (2003) showed that the Fourier-frequency resolved spectra can be reproduced in the propagation model, while Uttley (2004) argued for this model based on the r.m.s.--flux relation in accreting pulsar SAX J1808.4-3658 (see also Uttley \\& M$^{\\rm c}$Hardy 2001). ", "conclusions": "\\label{sec:discuss} We have computed the variability properties of the Fe \\Ka line from X--ray reprocessing in a propagation model of X--ray emission in accreting compact objects. The model combines results from various spectral and timing studies of accreting black holes. The former suggest the geometry of the standard optically thick accretion disc truncated at a radius larger than the radius of the last stable orbit (e.g.\\ Esin, McClintock \\& Narayan 1997; Gierli\\'{n}ski et al.\\ 1997; \\.{Z}ycki, Done \\& Smith 1998; Done \\& \\.{Z}ycki 1999; Zdziarski et al.\\ 1999, 2003). The latter postulate correlated flares (avalanches) and spectral evolution during flares, in order to explain power spectra and time lags (Poutanen \\& Fabian 1999; Kotov et al.\\ 2001). Combined spectral-timing studies suggest a connection between the geometry and timing properties (Revnivtsev et al.\\ 1999, 2001; \\.{Z}ycki 2002, 2003). Most of the observational results were obtained for black hole binaries, but, were possible to conduct, analysis of data for Seyfert galaxies confirm the general similarity between the two classes of objects (e.g.\\ spectral studies by Done et al.\\ 2000; Chiang et al.\\ 2000; Lubi\\'{n}ski \\& Zdziarski 2001; timing studies of Czerny et al.\\ 2001; Uttley \\& M$^{\\rm c}$Hardy 2001; Vaughan et al.\\ 2003; Markowitz et al.\\ 2003b; Papadakis 2004). \\begin{figure} \\epsfxsize = 8 cm \\epsfbox[18 250 620 710]{rmse.ps} \\caption{ Dependence of r.m.s.\\ variability on energy. Clear dip at Fe \\Ka line energy is seen, from the weakly variable line. \\label{fig:rmse}} \\end{figure} The variability properties of the Fe \\Ka line seem to be similar in both classes of objects in that the line appear to be less variable than the continuum which drives it, and, where the variability is detected, it does not seem to be clearly correlated with the continuum. This is contrary to simple(st) ideas, whereby the continuum and the line are produced in the same region and thus should be closely related. We note that the observational situation is far from clear, though. Time resolved spectral analysis would be the most direct method to determine the variability of the line, but the results may be model dependent (see detailed discussion in Zdziarski et al.\\ 2003). Decomposition of counts from a medium energy resolution instrument like RXTE/PCA into the line and continuum depends on the model assumed for both the line and the reflected continuum. In particular, the effects of relativistic smearing of the reflected continuum was not taken into account. We note though that according to Markowitz et al.\\ (2003a) the results on line variability in a sample of Seyfert galaxies were insensitive to assumptions about the amplitude of the reflected component: whether its relative amplitude was fixed, or allowed to vary in accord with the \\Ka line. Our model does reproduce the reduced variability of the Fe \\Ka line, compared to the variability of its driving continuum. Line variations may also appear not exactly correlated with continuum variations, because of the time delay between the peaks of the line and continuum fluxes. Both effects are necessary consequences of the adopted geometry, of a truncated disc with inner hot flow. The same geometry can also explain the hard X--ray time lags (Kotov et al.\\ 2001; \\.{Z}ycki 2003) through spectral evolution during flares (Poutanen \\& Fabian 1999). Quantitatively, our assumed ratio of heating to cooling rates, $C(r) \\propto \\lh(r)/\\lsoft(r)$, is a much weaker function of radius than could be expected for a compact (size $\\ll r$) active region and a sharply truncated disc. As already mentioned in \\.{Z}ycki (2003), in that latter geometry the supply of soft photons from the disc would diminish so rapidly that the predicted energy spectrum would be much too hard to be consistent with the data. This implies a {\\em gradual\\/} disappearance of the cold disc, which may be an interesting clue as to how the physical process of disc evaporation proceeds (R\\'{o}\\.{z}a\\'{n}ska \\& Czerny 2000). We emphasize that the reduction of variability of the \\Ka line discussed in the present paper is the same phenomenon as the decreasing reflection amplitude with Fourier frequency, found in X-ray data of Cyg X-1 and GX 339-4 by Revnivtsev et al.\\ (1999, 2001) and modelled by \\.{Z}ycki (2003). In the present paper this effect was analysed with tools usually applied to AGN data, which are usually analysed in time domain rather than Fourier domain. We note, that it does not seem possible to design the parameters of the model, so that the line flux remains exactly constant. This is because, even though the line may be constant during each flare (assuming $C(r) \\propto 1/\\lh(r)$), the number of flares active at any time varies, and this causes variation of the total flux of the line. A robust feature of the presented model is a connection of the time scale of the line {\\em response\\/} to the time scale of continuum variability. This is because the line flux responds on the time scale related to the duration of a flare. This in turn has to be chosen such that the observed power spectra are reproduced. The peaks in $f\\times P(f)$ are at rather long time-scales, much longer than just the light travel time through the region of most efficient energy generation. For example, the longest flares in our computations last $\\approx 10^6$ sec, which in light travel time corresponds to large distance of $2\\approx \\times 10^4\\,\\Rg$ (for $10^7\\,\\MSun$). This is simply a manifestation of the well known observational fact that time-scales of maximum X-ray variability power are much longer than the naively expected short dynamical time-scale. The physical mechanisms of producing the relatively long time scales of variability are rather unclear, but some interesting possibilities were recently considered in literature. 3-D magnetohydrodynamical simulations of Narayan, Igumenshchev \\& Abramowicz (2003) reveal slow ($v\\ll v_{\\rm ff}$) drift of plasma clumps across lines of magnetic field, which lines are compressed by the accretion flow. King et al.\\ (2004) considered a model involving magnetic dynamos operating locally in the disc. The longer and larger flares are results of correlations between dynamos acting in neighboring locations (radii). While individual dynamos operate of short (dynamical) time scales producing short flares, the correlated behaviour produces longer lasting, large events. Whatever the exact physical processes are, it is clear that the accretion flow is highly inhomogeneous and structured, factors that any realistic modelling should allow for. In the presented model, the line flux is leading the continuum flux. We ignored the light travel time delay of the line photons after the continuum, but this is unlikely to affect our result, since the line is supposed to originate close to the location of emission of primary radiation. The light travel time delay may be estimated as $\\delta t \\sim r\\times \\Rg/c$ (assuming the height of the emission region $h \\sim r$), which gives $\\delta t\\sim 10^3$ sec, i.e.\\ about the time bin in our simulations. Obviously, there may be additional effects due to, for example, adjustment of properties of the reprocessing medium to the increasing irradiation flux, which might affect the result to some extent. The dependence of r.m.s.\\ variability amplitude on energy is a relatively model independent demonstration of the reduced variability of the line. Our computations qualitatively reproduce the minimum of r.m.s.$(E)$ at the energy of the line. Generally, the r.m.s.\\ spectra are energy dependent variability amplitude, or, equivalently, can be thought of as representing energy spectra of the variable component of the spectrum. This can in principle be computed also in narrow ranges in Fourier frequency (Fourier frequency resolved spectroscopy, Revnivtsev et al.\\ 1999, 2001; \\.{Z}ycki 2002, 2003), but the quality of AGN data is not sufficient to make use of this technique as yet. \\begin{figure*} \\parbox{\\textwidth}{ \\parbox{0.45\\textwidth}{ \\epsfxsize = 0.4\\textwidth \\epsfbox[18 160 620 710]{flli4.ps} }\\hfil { \\parbox{0.45\\textwidth}{ \\epsfxsize = 0.4\\textwidth \\epsfbox[18 160 620 710]{fllifl4.ps} } }} \\caption{ Line flux vs.\\ the 7--30 keV flux. {\\it Left panel:\\/} each point shows one time bin of 5000 sec.\\ from a $10^7$ sec.\\ observation. The line flux follows linearly the continuum flux for low values of the latter, corresponding to the emission coming from outer regions (active region above the cold disc). The higher the continuum flux, the stronger the deviation from a linear relation, and the bigger the spread of points. {\\it Right panel:\\/} results from spectra binned in flux. The slope of the best fit line is $\\approx 0.56$. \\label{fig:flli}} \\end{figure*} Concluding, the propagation model of high energy emission in the geometry of a truncated accretion disc provides a framework for understanding many of the observed spectral and temporal characteristics of X-ray radiation from accreting black holes." }, "0403/astro-ph0403168_arXiv.txt": { "abstract": "We discuss the prompt emission of GRBs, allowing for $\\gamma\\gamma$ pair production and synchrotron self-absorption. The observed hard spectra suggest heavy pair-loading in GRBs. The re-emission of the generated pairs results in the energy transmission from high-energy gamma-rays to long-wavelength radiation. Due to strong self-absorption, the synchrotron radiation by pairs is in optically thick regime, showing a thermal-like spectral bump in the extreme-ultraviolet/soft X-ray band, other than the peak from the main burst. Recently, the prompt soft X-ray emission of GRB 031203 was detected thanks to the discovery of a delayed dust echo, and it seems to be consistent with the model prediction of a double-peak structure. The confirmation of the thermal-like feature and the double-peak structure by observation would indicate that the dominant radiation mechanism in GRBs is synchrotron rather than inverse-Compton radiation. ", "introduction": "In the past few years, a standard model was well established in which the gamma-ray burst (GRB) afterglows result from the relativistic blast-waves sweeping up the ambient medium of GRBs (\\Mesz 2002). However, the prompt emission of GRBs is believed to be irrelevant to ambient medium, and its radiation mechanism is still poorly known so far. The recent definite proof of GRB 030329 associated with a type Ib/c supernova confirmed, as long suspected, that GRBs, at less the long class, originate from explosions of massive stars in distant galaxies (Stanek \\etal 2003; Hjorth \\etal 2003). Since GRBs are events occurring on stars, the emission region may be compact, and the huge energy release will lead to the formation of $e^\\pm,\\gamma$ fireballs, exhibiting thermal-like spectra. But the GRB spectra are non-thermal and hard, with a significant fraction of the energy above the \\pair pair formation energy threshold. For a photon with tens of MeV to escape freely, avoiding $\\gamma\\gamma$ interactions, the fireball must be ultra-relativistic expanding, with $\\Gamma\\ga100$ (Lithwick \\& Sari 2001, and references therein). The afterglow studies has also confirmed the presence of ultra-relativistic motion. However, if the intrinsic emission, before leaking out from fireball, includes radiation with even higher energy, say, beyond GeV, these radiation still suffers $\\gamma\\gamma$ absorption, leading to pair loading in GRBs. In context of relativistic fireball model, Li \\etal (2003; Li03 hereafter) found that, in a wide range of model parameters, the resulting pairs may dominate those electrons associated with fireball baryons. The presence of abundant pairs would affect the behaviors of the early afterglow from reverse shocks (Li03), and may also emit particular signals in the bursting phase. We discuss in this Letter the prompt GRB emission, with emphasis on the re-emission by the secondary \\pair pairs. If the energy density in the emission region is dominated by magnetic field, the pairs would re-emit mainly by synchrotron radiation, rather than IC process (e.g., Pilla \\& Loeb 1998). Due to strong self-absorption, the pair emission appears as a thermal-like bump in the GRB spectrum, similar to the feature discussed by Kobayashi, \\Mesz \\& Zhang (2004) in the context of reverse shock emission. (Fan \\& Wei 2004 have also studied the pair emission, but with less stress on the self-absorption effect.) We further show that the intense soft X-ray emission in GRB 031203, inferred by the delayed dust halo (Vaughan \\etal 2004 [V04]; Watson \\etal 2004 [W04]), can be accounted for by the spectral bump due to pair-loading. This is of significant interest, since this feature could give a diagnostics for the magnetic field in the fireball (Kobayashi, \\Mesz \\& Zhang 2004) and the dominant radiation mechanism in GRBs. ", "conclusions": "We have studied the prompt GRB emission, allowing for $\\gamma\\gamma$ pair production and synchrotron self-absorption. Inferred by the observed characteristics of GRB emission, the resulting pairs usually dominate the baryonic electrons. The pairs will give rise to further emission by synchrotron radiation if in the strong magnetic field, which is also responsible to the prompt sub-MeV emission. However, due to strong self-absorption the pair emission exhibits a thermal-like bump in the extreme UV/soft X-ray band, other than the peak in the hard X-ray band. Since the feature emerges for $Y<1$, its observation gives a diagnostics for the magnetic energy density in the fireball (Kobayashi, \\Mesz, \\& Zhang 2004). The recent observation of a dust halo around GRB 031203 infers a spectral peak of the prompt burst emission in the soft X-ray band, which seems to be consistent with the predicted double-peak structure. Some primary hypotheses have been taken in our calculation. First, we assume that the emission region is transparent for Compton scattering, even though the secondary pairs increase significantly the total optical depth. For typical parameter values this assumption is protected. However, in some extreme cases with quite small $\\Gamma$ and $R$, the secondary pairs may form an optically thick screen again (Guetta, Spada \\& Waxman 2001; Kobayashi, Ryde \\& MacFadyen 2002), which degrades the gamma-rays and results in an X-ray flash (XRF; \\Mesz \\etal 2002). If so, our calculation using eq. (\\ref{Npm}) may underestimate the pair-loading in XRFs, which may need detailed works of numerical simulation (e.g., Pe'er \\& Waxman 2003). Secondly, we assume strong magnetic field, $Y<1$, in the emission region. If $Y>1$, the pairs lose most energy by IC scattering the GRB photons, and the IC photons are not self-absorbed again since beyond the optically thick regime, hence no effective energy exchange between pairs and photons is established and the bump disappears. Therefore, once UV/soft X-ray bumps are detected this will infer $Y<1$ and that it is synchrotron rather than IC radiation that gives rise to the sub-MeV emission of GRBs." }, "0403/astro-ph0403218_arXiv.txt": { "abstract": "We report the discovery of a locus of stars in the SDSS $g-r$ vs. $u-g$ color-color diagram that connects the colors of white dwarfs and M dwarfs. While its contrast with respect to the main stellar locus is only $\\sim$1:2300, this previously unrecognized feature includes 863 stars from the SDSS Data Release 1. The position and shape of the feature are in good agreement with predictions of a simple binary star model that consists of a white dwarf and an M dwarf, with the components' luminosity ratio controlling the position along this binary system locus. SDSS DR1 spectra for 47 of these objects strongly support this model. The absolute magnitude--color distribution inferred for the white dwarf component is in good agreement with the models of Bergeron et al. (1995). ", "introduction": "Modern large-scale accurate photometric surveys offer an unprecedented view of stellar populations. Here we discuss a population of unresolved binary stars which account for fewer than 10$^{-3}$ of stars detected by the Sloan Digital Sky Survey (York et al. 2000). Despite this low occurance frequency, the sample presented here is sufficiently large ($\\sim$1000 stars) to characterize their broad-band optical properties. \\subsection{ Sloan Digital Sky Survey} The Sloan Digital Sky Survey (SDSS; Abazajian et al.~2003, and references therein) is revolutionizing stellar astronomy by providing homogeneous and deep ($r < 22.5$) photometry in five passbands ($u$, $g$, $r$, $i$, and $z$; Fukugita et al. 1996, Gunn et al. 1998, Hogg et al. 2001, Smith et al. 2002), accurate to 0.02 mag (Ivezi\\'{c} et al.~2003). Ultimately, up to 10,000 deg$^2$ of sky in the Northern Galactic Cap will be surveyed. The survey sky coverage will result in photometric measurements for over 100 million stars and a similar number of galaxies. Astrometric positions are accurate to better than 0.1 arcsec per coordinate (rms) for point sources with $r<20.5^m$ (Pier et al.~2003), and the morphological information from the images allows robust star-galaxy separation to $r \\sim$ 21.5$^m$ (Lupton et al.~2003). Here we report the results of a color-based search for binary stars in the recent SDSS Data Release 1 (see www.sdss.org), which includes 53 million unique objects detected in 2099 deg$^2$ of sky. \\subsection{ The Stellar Locus in the SDSS Photometric System } The effective temperature is the dominant parameter that determines the position of the majority of stars in optical color-color diagrams constructed with broad-band filters (Lenz et al. 1998, and references therein). The effective temperature range results in a well-defined stellar locus in color-color diagrams (for more details see Finlator et al. 2000, and references therein). ", "conclusions": "The accurate multi-band SDSS photometry for a large number of stars allowed detection of a new feature in the broad-band optical color-color diagrams: a ``bridge'' of stars, well-separated from the main stellar locus, connects the positions of M dwarfs and white dwarfs. The bridge characteristics are consistent with a binary system than includes an M dwarf and a white dwarf, with the system's position on the bridge determined by the components' luminosity ratio. This conclusion is strongly supported by SDSS spectra for 47 such systems. The distance to these systems can be estimated in a straightforward way because a photometric parallax relation for M dwarfs can be applied to $i$ and $z$ band measurements, where the contribution from the white dwarf is negligible. With a known system distance, the white dwarf luminosity-color distributions can be determined and compared to models. We find that models by Bergeron et al. (1995) are in good agreement with the data. This work analyzed only about a quarter of the data that will be obtained by the SDSS. Thus, the color selection method presented here will eventually yield $\\sim$4,000 unresolved M dwarf-- white dwarf binary systems. \\vskip 0.3in {\\it Acknowledgements} We thank Princeton University for a generous support of this research. Funding for the creation and distribution of the SDSS Archive has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Aeronautics and Space Administration, the National Science Foundation, the U.S. Department of Energy, the Japanese Monbukagakusho, and the Max Planck Society. The SDSS Web site is http://www.sdss.org/." }, "0403/gr-qc0403069_arXiv.txt": { "abstract": " ", "introduction": "Thanks to the recent technological advance, there are many on-going projects of gravitational wave detection in the world; the Laser Interferometric Gravitational Wave Observatory (LIGO)~\\cite{LIGO}, VIRGO~\\cite{VIRGO}, GEO-600~\\cite{GEO}, ACIGA~\\cite{ACIGA}, TAMA300~\\cite{TAMA} and the Large-scale Cryogenic Gravitational wave Telescope (LCGT)~\\cite{LCGT} which are ground-based laser interferometers, and EXPLORER~\\cite{EXPLORER}, ALLEGRO~\\cite{ALLEGRO}, NIOBE~\\cite{NIOBE}, NAUTILUS~\\cite{NAUTILUS} and AURIGA~\\cite{AURIGA} which are bar detectors. Furthermore, there are some future space interferometer projects such as the Laser Interferometer Space Antenna (LISA)~\\cite{LISA} and the DECi hertz Interferometer Gravitational wave Observatory (DECIGO)~\\cite{DECIGO}. The detection of gravitational waves provides us with not only a direct experimental test of general relativity but also a new window to observe our Universe. To use them as a new tool of observation, it is necessary to derive theoretical waveforms. Once we know them, we may appeal to the matched filtering technique to extract source's information from gravitational wave signals. However, because the signals are expected to be very weak and the amount of data will be enormous for long-term continuous observations, it is essentially important to develop efficient data analysis methods. For these ground-based as well as future space-based interferometers, coalescence of compact object binaries is the most important source of gravitational waves. The process of a binary coalescence can be divided into three distinct phases. During the initial {\\it inspiral} phase, the gravitational radiation reaction timescale is much longer than the orbital period. The gravitational waves from the inspiral phase carry the information of the masses, spins and so on, of the system. After the inspiral phase, the binary becomes dynamically unstable and starts to merge. This phase is called the {\\it merger} phase. The gravitational waves from the merger phase give us the information about fully general relativistic dynamics of the system. Finally, if a black hole is formed after the merger phase, the system enters the {\\it ringdown} phase where it gradually settles down to a stationary Kerr black hole. During this process, the black hole emits gravitational waves with frequencies and damping rates specific to its mass and spin. Thus, the gravitational waves in the ringdown phase carry the information of the mass and spin of the final black hole. In this paper, we consider an effective search method for gravitational ringing of distorted spinning (Kerr) black holes. The ringdown waves are described by quasi-normal modes of a black hole. The quasi-normal modes are complex frequency wave solutions of the perturbed Einstein equations with purely outgoing-wave boundary condition at infinity and ingoing-wave at horizon, with vanishing incoming-wave amplitude. A quasi-normal mode is characterized by the central frequency $f_c$, usually called the (quasi-)normal-mode frequency, and the quality factor $Q$ which is inversely proportional to the imaginary part of the complex frequency. Their properties were analyzed extensively by Leaver~\\cite{Leaver}, and it is known that the least-damped (fundamental) mode belongs to the $\\ell=m=2$ spin-$2$ spheroidal harmonic modes. we note that a ringdown signal decays exponentially. Therefore, unless the signal-to-noise ratio (SNR) is very large, the signal will be soon buried in the noise after a few oscillation cycles. So, it is essentially important to keep the loss of SNR as small as possible when we construct a search method. Here we consider a black hole characterized by its mass $M$ and the dimensionless spin parameter $a=J/M^2$ where $J$ is the spin angular momentum. The parameter $a$ takes a value in the range $[0,1)$, with $a=0$ corresponding to a Schwarzschild black hole and $a=1$ to an extreme Kerr black hole. It is known that the imaginary part of the $\\ell=m=2$ least-damped mode is the smallest of all the quasi-normal modes, and results of black hole perturbation calculations as well as numerical relativity simulations strongly suggest that a ringdown wave is dominated by this $\\ell=m=2$ least-damped mode unless $a$ is extremely close to unity. Hence we focus on this single mode. Then a ringdown waveform is expressed as \\begin{eqnarray} h(f_c,\\,Q,\\,t_0,\\,\\phi_0;\\,t) = \\cases{ e^{ - \\frac {\\pi \\,f_c\\,(t-t_0)}{Q}}\\,\\cos(2\\,\\pi \\,f_c\\,(t-t_0)-\\phi_0) & for $t \\geq t_0$ \\,, \\cr 0 & for $t < t_0$ \\,, \\cr} \\label{eq:RDwave} \\end{eqnarray} where we set the amplitude to unity for simplicity, and $t_0$ and $\\phi_0$ are the initial time and phase of the ringdown wave, respectively. For the $\\ell=m=2$ least-damped mode, analytical fitting formulas for the central frequency $f_c$ and quality factor $Q$ were given by Echeverria~\\cite{ech} as \\begin{eqnarray} f_c &\\simeq& 32{\\rm kHz}\\,[1-0.63\\,(1-a)^{0.3}] \\left({M \\over M_{\\odot}}\\right)^{-1} \\,, \\label{eq:fcMa} \\\\ Q &\\simeq& 2.0\\,(1-a)^{-0.45} \\,. \\label{eq:QMa} \\end{eqnarray} There exists rich literature on search methods for ringdown waves. Echeverria~\\cite{ech} investigated the problem of extracting the black hole parameters from gravitational wave data in the case when SNR is large. Finn~\\cite{finn} improved this situation by developing a maximum likelihood analysis method that can deal with any SNR. Flanagan and Hughes then considered the parameter extraction from the three stages of a binary coalescence, i.e., from inspiral, merger and ringdown phases, in their series of papers~\\cite{FH}. For the ringdown phase, they discussed the relation between the energy spectrum of the radiation and SNR. Creighton~\\cite{crei} reported the result of analyzing data of the Caltech 40m by matched filtering, and emphasized the importance of coincidence event searches to discriminate spurious events from real events. But the search was limited to a single ringdown wave template. Recently, Arnaud et al.~\\cite{arna} discussed a tiling method to cover the 2-dimensional template space $\\{f_c,\\,Q\\}$. In our previous paper~\\cite{NTTS}, we proposed a more efficient method for tiling the template space. There, however, we ignored the dependence of the metric of the template space $\\{f_c,\\,Q\\}$ on the initial phase $\\phi_0$. This induces some small but non-negligible decrease in the match for a signal with certain ranges of $\\phi_0$. In this paper, we remove this shortcoming by properly taking into account the initial phase dependence, and develop a similar but substantially improved template spacing which is much more reliable than the previous one. Here we make a comment on an analysis using real interferometers' data. In this case, we have to deal with non-stationary, non-Gaussian noises, and Tsunesada and Kanda~\\cite{TsuneKanda} found that more fake events are observed than in the case of an inspiraling wave search, because the duration of a ringdown wave is typically much shorter than that of an inspiral wave and it can easily be affected by short bursts. It will be necessary develop a way to remove such fake events without losing real ringdown signals, but we leave this issue for future work. The paper is organized as follows. In Sec.~\\ref{sec:TDA}, first, we introduce orthonormal template waveforms for ringdown waves. Second, assuming a white noise background, we consider matched filtering in the 4-dimensional template space $\\{t_0,\\,\\phi_0,\\,f_c,\\,Q\\}$. We show that we can effectively reduce the template space to 2-dimensions spanned by $\\{f_c,Q\\}$, but the metric of this reduced template space depends on $\\phi_0$. Then, by carefully taking account of the $\\phi_0$ dependence of the metric, we analytically develop an efficient and reliable tiling method for ringdown wave searches. In Sec.~\\ref{sec:TN}, by using a fitting curve for the TAMA noise spectrum during the Data Taking 8 (DT8) in 2003, we show that our template spacing developed for white noise is valid even in the case of colored noise. Finally, Sec.~\\ref{sec:Dis} is devoted to summary and discussion. In Appendix~\\ref{app:TM}, we recapitulate the tiling method we proposed in~\\cite{NTTS}. In Appendix~\\ref{app:PEE}, we summarize the parameter estimation errors for ringdown signals by using the Fisher information matrix. ", "conclusions": "\\label{sec:Dis} The detection of ringdown waves is a direct confirmation of the existence of a black hole. The ringdown waves have damped sinusoidal waveforms that reflect the mass and spin of a black hole. In our previous paper~\\cite{NTTS}, we proposed an efficient method for tiling the templates for matched filtering in the 2-dimensional $\\{f_c,\\,Q\\}$ space. However, it relied on the template space metric for signals with the initial phase $\\phi_0=0$. In this paper, we took account of the initial phase dependence and developed a new, improved method to search for the ringdown waves. Since the template metric depends on the initial phase, we first determined the inner envelope curve of all the contours of a fixed maximum distance $ds^2_{\\rm max}$ between signals and templates on the $(f_c,\\,Q)$-plane for all possible values of the initial phase. We then constructed an ellipse that can cover the region inside the inner envelope as much as possible. Finally, we applied our tiling method proposed in~\\cite{NTTS} to obtain an improved, reliable tiling of the template space. Another change from our previous work is the difference in the definition of the match. In our previous work, we defined the match as the square of the SNR, $\\rho^2$, while here we used the SNR as the match. So, for the same fixed maximum distance, the number of templates needed to cover the template space turns out to be smaller than what we obtained previously, if we ignore the initial phase dependence. As a result, the actual number of templates to cover the template space did not increase much from what was suggested in~\\cite{NTTS}. We also examined the validity of our tiling method in the case of a colored noise spectrum by a Monte Carlo simulation. As a model of realistic noise power spectrum, we used a fitting curve of the noise power spectrum of TAMA300 during DT8 in 2003. For the pre-assigned maximum allowable SNR loss of $2\\%$, we found that only a few out of 2500 signals had the SNR loss larger than $2\\%$. This means that our template spacing is effective even in the case of colored noise. Since the real data includes the non-stationary, non-Gaussian noise, it is necessary to check the effectiveness of this method in a realistic situation. Using the real data of TAMA300, this new template spacing is now being tested by Tsunesada et al.~\\cite{Tsune}. They find a lot of fake events due to the non-stationary, non-Gaussian noise. Particularly, the detector has many noise sources that can produce fake ringdown wave signals. Apparently, we need to develop a method to remove these fake events without losing real gravitational ringdown wave signals. Perhaps, the best way is to perform a coincidence analysis if we have plural detectors. We plan to study methods of coincidence or coherent analyses by using several detectors in the future." }, "0403/astro-ph0403504_arXiv.txt": { "abstract": "{We explore the prospects for simultaneous, broad-band, multiwavelength polarimetric observations of GRB afterglows. We focus on the role of cosmic dust in GRB host galaxies on the observed percentage polarization of afterglows in the optical/near-infrared bands as a function of redshift. Our driving point is the afterglow of GRB 030329, for which we obtained polarimetric data in the $R$ band and $K$ band simultaneously $\\sim$1.5 days after the burst. We argue that polarimetric observations can be very sensitive to dust in a GRB host, because dust can render the polarization of an afterglow wavelength-dependent. We discuss the consequences for the interpretation of observational data and emphasize the important role of very early polarimetric follow-up observations in all bands, when afterglows are still bright, to study the physical properties of dust and magnetic fields in high-$z$ galaxies. ", "introduction": "Various independent studies of GRB afterglows have shown that in the optical bands the degree of linear polarization is usually rather low, in the order of 1 to 2\\% (for references see, e.g., Bj\\\"ornsson 2003; Covino et al. 2002, 2003; Greiner et al. 2003a; Lazzati et al. 2003; Masetti et al. 2003). If this turns out to be typical of optical afterglows (of long bursts), then the contribution of the dust in GRB host galaxies to the observed polarization properties of the afterglows may not be negligible. While dust in a GRB host galaxy will also manifest itself by changing the spectral energy distribution of the afterglow light away from an original power-law (e.g., Palazzi et al. 1998; Ramaprakash et al. 1998), polarization provides another independent method to constrain the existence of a dust component. If the sources responsible for the long bursts are located in star-forming regions, cosmic dust in the GRB environment will imprint a signal on the polarization properties of the afterglow light in the optical/near-infrared (NIR) bands. One might then imagine a situation in which the polarization by intervening dust (which might be time-independent) can dominate over the intrinsic polarization of the afterglow light. Motivated by our simultaneous optical and NIR polarimetric observations of the afterglow of GRB 030329, we explore here how dust in the GRB hosts may affect the polarimetry of the afterglow. Thereby, we concentrate on the wavelength dependence of the degree of linear polarization $P$. While the intrinsic polarization of the afterglow light is expected to be due to synchrotron radiation and, thus, wavelength independent, any additional polarization by dust in the GRB host can introduce a wavelength dependence to the measured $P$. Having this in mind, we basically focus on the redshift effect, which modifies this potential wavelength dependence of the degree of polarization (Serkowski 1973; Serkowski et al. 1975). Recently, Lazzati et al. (2003) have discussed the potential imprint of cosmic dust in GRB hosts on the interpretation of polarization data, with particular emphasis on the afterglow of GRB 021004. Whereas these authors concentrated on the interpretation of the time evolution of the percentage polarization, $P(t)$, here we focus on the wavelength dependence of $P$ at a fixed time. The latter comes into play if dust dominates the polarimetric properties of the afterglow light. While the time evolution of $P$ is basically an indicator for the GRB outflow geometry (Ghisellini \\& Lazzati 1999; Sari 1999; Rossi et al. 2004; Lazzati et al. 2004; for the cannonball model see Dado et al. 2004), the wavelength dependence of $P$ is a signature for the presence of dust in the GRB host. Also, whereas a measure of $P(t)$ requires an intense monitoring of the rapidly fading afterglow over many days, the wavelength dependence of $P$ can be investigated at any time, in particular when the afterglow is still in its earliest evolutionary phase, i.e., when the afterglow is brightest. Rapid multiwavelength polarimetric observations of GRB afterglows could then allow us to correct for the influence of dust once the time evolution of $P$ is studied. ", "conclusions": "Encouraged by our successful simultaneous optical and NIR polarimetric observations of the afterglow of GRB 030329, we have explored here in which manner simultaneous multicolor polarimetric observations can reveal an imprint from cosmic dust in GRB host galaxies on the afterglow light. We have argued that dust in GRB hosts can strongly affect the polarization properties of an afterglow, depending on its redshift and on the photometric band, in which the observations are performed. In particular, the wavelength dependence of polarization due to dust, which is in strong contrast to the intrinsic wavelength independence of the fireball light, could be used as an indicator for the presence of dust. With the upcoming $Swift$ satellite mission GRB afterglows are expected to be localized within seconds after burst trigger, in principle allowing for first polarimetric observations within the first hour after a burst. Since the amount of maximum polarization, $P_{\\rm max}$, can rapidly increase with increasing visual extinction within the host galaxy (Schmidt-Kaler 1958; Serkowski et al. 1975), in principle large degrees of linear polarization could be detectable for the most intrinsically extinguished afterglows. Conversely, if a wavelength dependent polarization is never found in an afterglow then either the host dust properties must be substantially different from those of the Galactic ISM, or there is no dust at all, or the magnetic field in the host galaxy is weak and/or not well ordered. The position angle of the polarization defines the orientation of the predominant magnetic field in the host galaxy along the line of sight. This predominant field can either be the large scale magnetic field acting in the GRB host (in particular, if the host is seen edge-on), or the field characterizing the star-forming region in which the burster is placed. The detection of interstellar dust extinction in the host without a substantial polarization signal could hint to a strongly tangled magnetic field structure in the host galaxy, or even indicate the absence of a magnetic field large enough ($\\sim 10^{-8}$~G) to align the dust grains in the host galaxy. Both effects would indicate very strong evolutionary effects in the galactic magnetic field structure as practically all well studied nearby galaxies are known to have ordered magnetic field on large ($\\sim$~kpc) scales (Kronberg 1994; Scarrott 1996; Wielebinski \\& Krause 1993). In the Milky Way the dust grain alignment along the arbitrary line of sights due to the Galactic magnetic field produces a polarization per magnitude of visual extinction in the range $P/A_{\\rm V}\\approx 0.6-3~\\%\\,{\\rm mag}^{-1}$ (Serkowski et al. 1995). Thus a polarization below this value might indicate special conditions in a GRB host galaxy, including the star-forming region in which the burster is placed." }, "0403/hep-th0403075_arXiv.txt": { "abstract": "We derive the $4D$ low energy effective field theory for a closed string gas on a time dependent FRW background. We examine the solutions and find that although the Brandenberger-Vafa mechanism at late times no longer leads to radion stabilization, the radion rolls slowly enough that the scenario is still of interest. In particular, we find a simple example of the string inspired dark matter recently proposed by Gubser and Peebles. ", "introduction": "String theory continues to be our leading candidate for a quantum theory of gravity. However, it continues to be a challenge to find phenomenological predictions that can be verified by experiment. One challenge that stymies the effort towards string phenomenology is our lack of understanding of the ground state of string theory. This has come to be known as the cosmological moduli problem (see e.g. \\cite{dine} and references within). One approach to resolving this problem is to take the point of view that cosmological evolution should be responsible for determining the value of the moduli. One often finds that, by adopting this view, the moduli relax to special locations in the stringy landscape, which are points of enhanced symmetry. A recent example of such a scenario was presented in \\cite{stanford}. There it was found that the moduli are trapped in orbits around points of enhanced symmetry due to the production of light string modes. Stability then sets in due to the Hubble damping resulting from cosmological evolution. Alternatively, one can also have so-called racetrack models where the moduli continue to roll and do not remain fixed, but could give interesting cosmological consequences \\cite{quevedo}. In this latter approach an important issue is fine tuning or the cosmic coincidence problem, i.e. why did the moduli start rolling at such a particular time? The Brandenberger-Vafa scenario (BV scenario) \\cite{bv}, or Brane Gas Cosmology (BGC) \\cite{loitering,branes,stable,isotropy,perturbations,columbia,columbia2,subodh} as it has come to be known, is an example of a cosmological model that incorporates both of these approaches to the moduli problem. In these scenarios one generally works in the background of ten dimensional dilaton gravity\\footnote{Although this has been extended to M-theoretical considerations in \\cite{columbia2}.}, with sources given by the string winding and momentum modes. It is found that the scale of the extra dimensions (radion) is then stabilized at the self-dual radius, where many of the string modes become massless and the symmetry is enhanced \\cite{stable}\\footnote{We note that the analysis of \\cite{stable} was performed with the case of bosonic strings in mind. Stabilization at the self-dual radius leading to enhanced gauge symmetry is expected for heterotic strings, but this does not occur for Type II strings. We thank Steve Gubser for reminding us of this point.}. A crucial aspect of these findings was the running of the dilaton to weak coupling, which was driven by the winding and momentum modes of the string. In addition, it has been shown that this model is stable to both anisotropies \\cite{isotropy} and inhomogeneities at the linear level \\cite{perturbations}. In this paper we would like to extend the BV scenario to better understand its predictions for late-time cosmology. In particular, we would like to see if the stabilization mechanism is still plausible in the $4D$ effective field theory resulting from dimensional reduction. Here, one usually assumes that the dilaton is fixed, since otherwise this would lead to unacceptable observational consequences\\footnote{However there have been attempts to keep the dilaton dynamical, see e.g. \\cite{polyakov}}. Given that at late times we are no longer in the regime of dilaton gravity, perhaps one would naively expect that such stabilization would no longer work. That is, in General Relativity one must generally introduce exotic matter and/or violate the weak energy condition to stabilize extra dimensions \\cite{Carroll}. The string modes that we will consider here do not have such properties. To spare the reader suspense, we do in-fact find that stabilization fails, except in the special case of one extra dimension\\footnote{We mention that the case of $d=1$, from the perspective of the $5D$ Einstein frame, has already been considered in \\cite{subodh}.}. However, this is not as disastrous as one might first imagine. In fact, although the radion is no longer stable in the effective theory, its evolution is slow enough compared to that of the $4D$ background to be observationally acceptable. In addition, we find that this evolution can lead to interesting phenomenology. As an example, we find an example of a cold dark matter candidate like that recently purposed by Gubser and Peebles \\cite{gubser}. In Section II, we will briefly review the radion stabilization mechanism as presented in \\cite{stable}. We present the stress energy tensor for the string modes and the corresponding action for the string modes. In Section III, we consider the evolution in the Einstein frame. This is not the correct frame at early times when one is interested in the geodesics followed by the strings, but will be important for the late-time cosmology. In Section IV, we dimensionally reduce the theory and find a form for the string modes in the $4D$ effective theory, which is given by their effective potential. From the effective potential we are able to discuss the stability of the radion, which is shown to exhibit slow rolling behavior. This leads us to the possibility of closed strings in the extra dimensions behaving as dark matter. ", "conclusions": "In this article we have shown that stabilization of the radion in the effective theory at late-times is problematic. This was not the case with one extra dimension, where after passing to the effective theory the potential retains a minimum at the self-dual radius. In that case, we were able to construct a simple realization of scalar dark matter as discussed in \\cite{gubser}. However, in the case of more than one extra dimension the potential does not possess a local minimum. This led to extra dimensions that are slowly growing, but that remain small compared to the large dimensions. This case is interesting in many respects, e.g. it could lead to {\\em large} extra dimensions without the need to invoke brane world scenarios. We also note that these results remain valid for both a fixed and evolving dilaton, where in the later case the string modes drive the dilaton towards the region of weak coupling. We find the interpretation of winding and momentum modes as CDM much more involved, but possible. We leave such considerations for future work." }, "0403/astro-ph0403075_arXiv.txt": { "abstract": "Several recent papers have studied lensing of the CMB by large-scale structures, which probes the projected matter distribution from $z=10^3$ to $z\\simeq 0$. This interest is motivated in part by upcoming high resolution, high sensitivity CMB experiments, such as APEX/SZ, ACT, SPT or Planck, which should be sensitive to lensing. In this paper we examine the reconstruction of the large-scale dark matter distribution from lensed CMB temperature anisotropies. We go beyond previous work in using numerical simulations to include higher order, non-Gaussian effects and find that the convergence and its power spectrum are biased, with the bias increasing with the angular resolution. We also study the contamination by the kinetic Sunyaev-Zel'dovich signal, which is spectrally indistinguishable from lensed CMB anisotropies, and find that it leads to an overestimate of the convergence. We finish by estimating the sensitivity of the previously cited experiments and find that all of them could detect the lensing effect, but would be biased at around the 10\\% level. ", "introduction": "\\label{sec:intro} Weak gravitational lensing by large-scale structure has recently become a powerful tool in the cosmologists' toolbox, allowing us to map the mass distribution in the universe. Lensing measurements using galaxy ellipticities have already begun to constrain cosmological parameters and to test our paradigm for hierarchical structure formation \\citep[see][for a recent review]{vWMe03}. Even as this effort ramps up, yet another doorway into the weak lensing gold mine will open as a new breed of large surveys, such as the Atacama Cosmology Telescope (ACT\\footnote{http://www.hep.upenn.edu/$\\sim$angelica/act/act.html}), APEX-SZ\\footnote{http://bolo.berkeley.edu/apexsz/}, Planck\\footnote{http://astro.estec.esa.nl/Planck}, and the South Pole Telescope (SPT\\footnote{http://astro.uchicago.edu/spt/}), begin to come on line. These surveys will probe the millimeter and sub-millimeter wavebands with unprecedented power and resolution, and thus enable us to map the large-scale distribution of mass in the universe by probing the gravity induced deflections of cosmic microwave background (CMB) photons as they journey from primordial times at $z\\sim 10^3$ to the present. In addition to providing the first ever map of the projected dark matter over most of cosmological history, such a measurement may have the potential to provide us with precision constraints on cosmological parameters, such as the neutrino mass and the dark energy equation of state \\citep{Kaplinghat03}, by measuring the matter power spectrum with percent level accuracy.\\\\\\\\ Beginning with the pioneering work of \\cite{Zaldarriaga99}, a considerable effort has been put into developing an accurate and sensitive estimator of the lensing effect. In this paper, we seek to further the effort to develop methods necessary for the best possible estimation of the projected mass distribution from CMB temperature information. To this end, we examine several effects which have not been studied in earlier works, both to see how they influence the reconstruction and whether they will provide us an opportunity to enhance the signal-to-noise of the estimator. We note that although we will concentrate here on the extraction of information on large angular scales, reconstruction at the cluster level \\citep{Seljak00} is another exciting possibility, which we discuss elsewhere \\citep{VAW04}. The current state of the art techniques comprise maximum likelihood estimators \\citep{Hirata03} and the computationally more tractable quadratic estimators \\citep{Hu01a}. These assume that both the primary CMB and the large scale structure responsible for the lensing are Gaussian random fields, and that noise is both Gaussian and uncorrelated with the signal. The estimators are optimized to solve for the lensing effect by looking for the non-Gaussianity induced by the mapping of one Gaussian field by another. Unfortunately, only the first of the four assumptions listed above is true. The projected mass distribution is non-Gaussian except on large angular scales, and the kinetic Sunyaev-Zel'dovich (kSZ) effect \\citep[][for recent reviews see \\citealt{Reph95,Birk99,Carl02}]{SZ72,SZ80a} is a particularly pernicious source of confusion because it is non-Gaussian, spatially correlated with many of the structures doing the lensing, and spectrally indistinguishable from primary CMB anisotropies. To test the impact of these issues, we create lensing and kSZ fields by using an N-body simulation to model relevant structures, and then apply these fields to random realizations of the CMB (details of our methods are provided in an Appendix). The maps that result from this are then used to reconstruct the projected mass density and the dark matter power spectrum for various experimental parameters. We have elected to use the quadratic estimator of \\cite{Hu01a} for our reconstructions, in part because it is computationally more tractable than the maximum likelihood estimator, and because the two methods are predicted to be roughly equivalent for the angular scales and experimental parameters we are considering \\citep{Hirata03}. More important, we hope to test the validity of the \\emph{assumptions} of the current generation of estimators, and thereby gain insight into the issues facing any method which is based on a Gaussian approximation.\\\\\\\\ We note that while our ability to remove foregrounds and point-sources from the CMB maps may ultimately prove challenging for any reconstruction, we will not focus on these complications here. Instead, we will restrict our analysis to the observationally irreducible complications discussed above and instrument effects which are roughly consistent with our fiducial surveys. Also, we will consider only the CMB \\emph{temperature} anisotropies, both because this simplifies the calculations and because it is more relevant for the next generation of wide field instruments with high angular resolution, which will not initially be polarization sensitive. This subject remains of interest, however, and certainly merits continued study. \\\\\\\\ The outline of our paper is as follows. In Section \\ref{sec:estimator} we describe the estimator of the projected mass distribution used hereafter, and motivate our choice. We then use this estimator to reconstruct a Gaussian projected mass map in Section \\ref{sec:gauss}. In Section \\ref{sec:nongauss} we investigate the effect of non-Gaussianity in the lensing field, and in Section \\ref{sec:ksz} we include contamination from the kSZ and describe our efforts to mitigate the problem. The effects of these contaminants are considered as a function of instrument resolution in Section \\ref{sec:surveys}, where we also provide an estimate of how well our fiducial surveys will reconstruct convergence maps and power spectra. We then summarize and discuss our results in Section \\ref{sec:summary}. Finally, some details of the simulations and the estimator are presented in an Appendix. ", "conclusions": "\\label{sec:summary} Lensing of CMB photons may provide a window into the matter distribution projected from primordial times to the present. A great deal of work has been put into developing the statistical methods needed to fulfill this promise, and our goal here has been to further these efforts. To this end, we have extensively checked the power and assumptions of one of the two principal estimators that have emerged, the ``optimal'' quadratic estimator of \\cite{Hu01a}, both for reconstructing projected mass maps and for measuring the matter power spectrum. Although our results apply specifically to the quadratic estimator, we note that the other principal candidate \\citep[the maximum likelihood estimator of][]{Hirata03} makes the same assumptions about the underlying fields as the quadratic estimator, so it is likely to face the same challenges. We began our efforts in Section \\ref{sec:gauss}, where we tested the quadratic estimator method under the artificial conditions of a Gaussian gravitational potential, and uncorrelated Gaussian noise, for which the estimator is most likely to succeed. The noise in the reconstructed maps has $\\kappa$ dependent extra terms which constitute an additive bias in power spectrum measurements. The second order terms were discussed in \\citep{Cooray03}, and we confirm their prediction that these are numerically significant. In addition, we find for the first time that higher order terms are also likely to be significant, since these second order terms do not fully correct bias for interesting observational parameters. Although the bias is signal dependent, it may be possible to correct for this fact by using iterative methods or model fits, as suggested by \\cite{Kesden03}; however, high order terms would need to be included to obtain good accuracy. It is clear that a reconstruction method based on the assumption of a Gaussian lensing field will not optimally treat non-Gaussian features, such as clusters, which contain more information than just the 2-point function. We investigated this issue in Section \\ref{sec:nongauss} by using a lensing field generated from an N-body simulation. This imposed both a multiplicative bias on the cross spectrum estimate, and an additional additive bias on the auto spectrum. The bias increased as the signal-to-noise ratio increased. We investigated the impact of a second non-Gaussian complication, the kinetic Sunyaev-Zel'dovich effect (kSZ), in Section \\ref{sec:ksz}. This ``noise'' is highly correlated with the lensing signal and, unlike the larger thermal SZ, is impossible to distinguish from the CMB using spectroscopic methods. We began by showing that, if left untreated, this effect will completely dominate the reconstruction. One obvious method to correct for this is to use a measurement of the thermal SZ (which is correlated with the kSZ) to mask pixels which are likely to have a large contamination, and we were able to substantially reduce the bias caused by the kSZ while excising only a relatively small number of pixels. Although this was encouraging, not all of the kSZ signal is correlated with the thermal SZ and the contamination level remained significant. Further reduction of kSZ noise by simply masking more pixels suffers from steadily decreasing returns in signal-to-noise improvement. We suspect, although we have not checked it explicitly, that contamination {}from the Ostriker-Vishniac effect would behave in a similar manner. Given the difficulty that one encounters using Gaussian reconstruction methods, it is natural to ask if the non-Gaussian nature of the lensing signal might itself provide the best probe of the lensing field. This topic has recently been the subject of great interest in the context of galaxy lensing \\citep[e.g.][]{TaJa03,ZaSc03,HoWh03}, but to this point a similar effort is lacking for lensing of the CMB. There is clearly promise in this idea; however, any measurement of the non-Gaussian signal is likely to be particularly sensitive to the masking technique needed to control the kSZ, which will be highly correlated with this signal. For example, clusters will produce both a large non-Gaussian lensing signal and a large kSZ in the same location, \\emph{at the cluster}. Thus, it may be necessary to mask precisely those pixels which would otherwise have provided the best window into the non-Gaussian signal. In this paper, we have been primarily interested in the reconstructions that can be accomplished using the unprecedented power and resolution that will be available in the next generation of surveys, and it is in this context that biasing effects emerge as important obstacles. In the regime of large smoothing and low signal-to-noise, the bias due to non-Gaussian effects is small, and the quadratic estimator performs well in this regime. It is only when the observational parameters are improved that the biasing effects we have been discussing emerge. We show in Fig. \\ref{fig:biasres} the unfortunate increase in bias that occurs as signal-to-noise is improved, so that a recovery of the power spectrum can have either small error bars or negligible bias, but not both. Thus, while a detection of the lensing effect is likely to be achieved by APEX/SZ, more ambitious goals that rely on extremely high fidelity reconstructions of the matter power spectrum require significantly more modeling. We have presented a number of reconstructed maps in this paper, and it is clear that features in these maps do coincide with the those of the original fields at some level. However, we note that the maps have been smoothed by a $20^\\prime$ FWHM Gaussian window, and even so the reconstructed maps are far from perfect. We have also performed extensive numerical integrations to compute the corrections to the noise term $\\mathcal{N}_{\\ell}$ outlined in Section \\ref{sec:gauss}. Although these corrections proved useful for for removing some of the bias effects, the computational power needed for this calculation is non-trivial, a point which we discuss in more detail in the Appendix. Although we have not treated the issue here, we would like to make some comments about the role of polarization in CMB lensing. It has been emphasized \\citep{Guzik00,Hu02,Cooray03,Hirata03b,Kesden03,Okamoto03} that the inclusion of polarization information might dramatically enhance the prospects for large-scale structure reconstruction from lensing of the CMB. This is because lensing induces a $B$-mode polarization signal which is otherwise absent for purely scalar, primary fluctuations. The large intrinsic primary CMB anisotropies, which are a source of ``noise'' for lensing reconstruction, are thus absent. However, the spatial structure is complicated for polarization, as it is for temperature, and the signal levels are much smaller. The kSZ effect, which is one of our major contaminants, is also polarized \\citep{SZ80b}. Thus, although it is certainly reasonable that the addition of polarization information would enhance the prospects for $C_\\ell^{\\kappa\\kappa}$ reconstruction, a detailed calculation is required to determine how much better one can actually do. Weak lensing is one of the principal tools for cosmologists, and promises to be increasingly significant for many years to come. Lensing of the CMB is a new addition to the arsenal, which will be facilitated by powerful upcoming surveys such as ACT, APEX-SZ, Planck, and SPT. Even as observers gear up to probe the millimeter and sub-millimeter wavebands with unprecedented power and resolution over large fractions of the sky, theorists continue to improve the methods necessary to extract relevant information from the resulting measurements. If the pace of development on both fronts continues at its current rate, the future looks bright indeed. \\noindent {\\bf Acknowledgments:}\\newline A.A. would like to thank the organizers of the workshop ``Cosmology with Sunyaev-Zel'dovich cluster surveys'' held in Chicago in September 2003 for allowing him the chance to present some of this work. Additionally we would like to thank T. Chang, J. Cohn, M. Zaldarriaga for helpful discussions about these results. The simulations used here were performed on the IBM-SP2 at the National Energy Research Scientific Computing Center. This research was supported by the NSF and NASA. \\appendix" }, "0403/astro-ph0403596_arXiv.txt": { "abstract": "{ We present an \\xmm mosaic observation of the hot ($kT\\sim6.5$~keV) and nearby ($z=0.0881$) relaxed cluster of galaxies A478. We derive precise gas density, gas temperature, gas mass and total mass profiles up to $12\\arcmin$ (about half of the virial radius $R_{200}$). The gas density profile is highly peaked towards the center and the surface brightness profile is well fitted by a sum of three $\\beta$--models. The derived gas density profile is in excellent agreement, both in shape and in normalization, with the published Chandra density profile (measured within $5\\arcmin$ of the center). Projection and PSF effects on the temperature profile determination are thoroughly investigated. The derived radial temperature structure is as expected for a cluster hosting a cooling core, with a strong negative gradient at the cluster center. The temperature rises from $\\sim2$~keV up to a plateau of $\\sim6.5$~keV beyond $2\\arcmin$ (i.e. $r>208\\rm{kpc}=0.1\\, R_{200}$, $R_{200}=2.08$~Mpc being the virial radius). From the temperature profile and the density profile and on the hypothesis of hydrostatic equilibrium, we derived the total mass profile of A478 down to 0.01 and up to 0.5 times the virial radius. We tested different dark matter models against the observed mass profile. The Navarro, Frenk \\& White (\\cite{navarro97}) model is significantly preferred to other models. It leads to a total mass of $M_{200}=1.1\\times 10^{15}$~M$_\\odot$ for a concentration parameter of $c=4.2\\pm0.4$. The gas mass fraction increases slightly with radius. The gas mass fraction at a density contrast of $\\delta=2500$ is $\\fgas=0.13\\pm0.02$, consistent with previous results on similar hot and massive clusters. We confirm the excess of absorption in the direction of A478. The derived absorbing column density exceeds the 21~cm measurement by a factor of $\\sim2$, this excess extending well beyond the cool core region. Through the study of this absorbing component and a cross correlation with infrared data, we argue that the absorption excess is of Galactic origin, rather than intrinsic to the cluster. ", "introduction": "As nodes of large scale structure and thus places of dark matter concentration, galaxy clusters can be used as powerful tools to test theories of structure formation. The basic hierarchical scenarios based on gravitation make the population of galaxy clusters a homologous population of sources. Their physical properties follow scaling laws depending only on their redshift and mass, and their internal structures are similar. The exceptional capabilities of \\xmm in terms of sensitivity and of Chandra in term of spatial resolution allow us to characterize the gas density and temperature profiles with unprecedented accuracy. For a relaxed cluster, the hydrostatic equations can be used to derive the underlying dark matter distribution, from the very central part of clusters up to nearly the virial radius (David \\etal~\\cite{david01}; Allen \\etal~\\cite{allen01b}; Arabadjis, Bautz \\& Garmire~\\cite{abg02}; Allen, Schmidt \\& Fabian~\\cite{allen02a}; Pratt \\& Arnaud~\\cite{pratt02},\\cite{pratt03}; Lewis \\etal~\\cite{lewis03}; Buote \\& Lewis~\\cite{buote03}). The observed clusters seem to have a cusped dark matter profile as predicted by numerical simulations (Navarro, Frenk \\& White~\\cite{navarro97}; hereafter NFW; Moore \\etal~\\cite{moore99}; hereafter MQGSL). However, the central slope of the dark matter profile and the possible dispersion of the concentration parameter remain open issues. Larger samples of high quality mass profiles are needed to further assess these points. In this paper we present the \\xmm spectro-imaging observation of A478, a massive, relaxed nearby cluster ($z=0.0881$ -- \\cite{struble99}). Detected in surveys (UHURU, HEAO-1, Ariel-V), this cluster is well known in X-rays and its physical properties have been carefully studied with previous X-ray observatories: EXOSAT (Edge \\& Stewart~\\cite{edge91}), Einstein and Ginga (Johnstone \\etal~\\cite{johnstone92}), ROSAT (Allen \\etal~\\cite{allen93}; White \\etal~\\cite{white94}) and ASCA (\\cite{markevitch98}, White \\etal \\cite{white00}). All those previous studies converge for what concerns the overall temperature of the cluster, $kT \\sim 6.8$~keV. Recently, Sun \\etal (\\cite{sun03}) performed a high angular resolution study of the central part of the cluster with Chandra. They pointed out the presence of an X-ray cavity in the very central part of the cluster which is anti-correlated with the radio lobes. Here we focus on the characterization of the gas and dark matter distribution of A478. In a companion paper de Plaa \\etal (in prep.) present a detailed spectroscopic study of the metal abundances and their distribution within A478's core based on EPIC and RGS data. We present the observation and the different data processing steps in Sect.~\\ref{sec:data}. In Sect.~\\ref{sec:morpho} we briefly discuss the cluster morphology. In Sect.~\\ref{sec:ne} we analyze the surface brightness profile and derive the gas density profile. Spatially resolved spectroscopic analysis is presented in Sect.~\\ref{sec:spec}, where we also discuss the temperature and absorption profiles. In Sect.~\\ref{sec:mass}, we present the resulting total mass and gas mass fraction profiles of A478 and we discuss the shape of the dark matter profile according to our observational results. Throughout this paper, we use $H_0=70$~km Mpc$^{-1}$ s$^{-1}$, $\\Omega_m=0.3$ and $\\Omega_{\\Lambda}=0.7$. In such a cosmological framework, at the cluster redshift ($z=0.0881$) $1$\\arcmin~$=99$~kpc. ", "conclusions": "In this paper we have analysed imaging and spatially resolved spectral data of the galaxy cluster A478 obtained with the \\xmm satellite. We obtained well constrained absorption, gas density and gas temperature profiles up to $\\sim 0.5$ virial radius. As in previous studies, we found an excess of absorption in the direction of A478. The derived absorbing column density exceeds the 21~cm measurements by a factor of $\\sim2$ in the center and the excess extends well beyond the cooling core region. This excess of absorption seen in A478 (and other cooling flow clusters) was interpreted in pre-XMM and Chandra studies (e.g. Allen \\etal \\cite{allen93,fabian94}) as the signature of intrinsic cool absorbing material, a consequence of the strong cooling flow in the cluster center. From the absorption excess extent and a detailed comparison with FIR data, we argue that the absorption excess is rather of Galactic origin. We suggest it could be the effect of a Galactic molecular/cold cloud type structure in the line of sight. The next generation of FIR space missions will help to clarify this issue with sensitive FIR mapping of the whole cluster area with a high spatial resolution. We fitted the surface brightness profile with various parametric models of the gas density profile, taking into account the \\xmm PSF. The gas density profile, derived on scales of $0.03\\arcmin - 13\\arcmin$, is highly peaked towards the center and is well fitted by a quadratic sum of three \\betamod. The derived gas density profile is in excellent agreement both in shape and normalization with the Chandra density profile (measured up to $5\\arcmin$ of the center). This indicates that the PSF modeling we have used is basically correct and that accurate density profiles in the very center of the cluster can be derived with XMM data, in spite of the PSF blurring. A raw temperature profile was obtained on scales of $0.07\\arcmin - 10\\arcmin$ by fitting isothermal models to spectra extracted in $13$ concentric annuli. This profile shows a sharp negative gradient measured toward the center ($r<2\\arcmin$), a signature of a cooling core. Beyond that region the profile is essentially flat. We have thoroughly investigated projection and PSF effects on the temperature profile determination. The PSF effects beyond $0.3\\arcmin$ are much less important than projection effects, whereas both are important in the very center. We discuss the noise introduced by the correction of these effects and a way to overcome this problem. The derived deprojected PSF-corrected temperature profile ranges from $\\sim2$~keV in the center up to an asymptotic value of $\\sim6.5$~keV. Using this temperature profile and the density profile, we have derived the total mass profile for this cluster from $0.01$ up to $\\sim0.5$ times the virial radius. Systematic uncertainties due to the PSF and projection correction for the temperature profile are taken into account. We have tested different models for the dark matter profile distribution against the observed mass profile. A mass distribution with a cusp in the center, as predicted from numerical simulations, is clearly preferred. An isothermal sphere model does not provide a good fit to the data. In a second time we tested an MQGSL model and an NFW model. Those two types of models have similar shapes at large radii (they both scale like $r^{-1.5}$) but differ significantly at small radii. Therefore to discriminate those two models one needs data with a high statistic quality over a wide range of radii (i.e. covering at least two decades). Our data set nearly fulfilled this requirement and we were indeed able to discriminate between the two models, the NFW model being preferred to the MQGSL model. For the NFW model, we derived a concentration parameter $c=4.2\\pm 0.4$. This value is as expected from numerical simulations: $c \\sim6$ (Navarro, Frenk \\& White \\cite{navarro97}; \\cite{eke98}) with a typical $1\\sigma$ dispersion of $\\Delta(\\log(c))=0.18$ (Bullock \\etal~\\cite{bullock} ). This work can be compared with the similar work on the cluster A1413 by Pratt \\& Arnaud (\\cite{pratt02}). In the case of A1413, if the NFW model was acceptable, the MQGSL model was slightly preferred. Although this cluster is detected out to 0.7 times the virial radius, the data are limited in the center, a shortcoming, as emphasized above, for discriminating between those two models. Moreover, data in the center only are not sufficient (see the work on A1983 by Pratt \\& Arnaud \\cite{pratt03}). On the other hand, our result agrees with the analysis of A2029 by Lewis et al. (\\cite{lewis03}), which clearly favors an NFW dark matter profile. To our knowledge, this is the only other data set which covers a similar wide radial range (0.001 to 0.1 virial radius) . The key factors in fitting the mass profiles with different dark matter models, are the resolution in the center as well as the data at large radii. To date \\xmm is the best satellite to compute total mass profiles, especially through its capability to derive precise temperature profiles. Nevertheless, its spatial resolution limits the investigation at the very center of galaxy clusters. A direct combined analysis of \\xmm and Chandra data of very well relaxed clusters seems to be an ideal path to a full description of the dark matter profile in clusters. However, one has to keep in mind that this requires an excellent cross calibration between the two satellites, so that the temperature profiles derived at various scales can be combined." }, "0403/astro-ph0403243_arXiv.txt": { "abstract": "An empirical relation between the broad line region (BLR) size and optical continuum luminosity is often adopted to estimate the BLR size and then the black hole mass of AGNs. However, optical luminosity may not be a good indicator of photoionizing luminosity for extremely radio-loud AGNs because the jets usually contribute significantly to the optical continuum. Therefore, the black hole masses derived for blazar-type AGNs with this method are probably overestimated. Here we first derived a tight empirical relation between the BLR size and the H$_\\beta$ emission line luminosity, $R(\\rm{light-days})= 24.05(L_{H_\\beta}/10^{42} ergs~s^{-1})^{0.68}$, from a sample of 34 AGNs with the BLR size estimated with the reverberation mapping technique. Then we applied this relation to estimate the black hole masses of some AGNs and found that for many extremely radio-loud AGNs the black hole masses obtained with the $R-L_{H_\\beta}$ relation are systematically lower than those derived previously with the $R-L_{5100\\AA}$ relation, while for radio-quiet and slightly radio-loud AGNs the results obtained with these two methods are almost the same. The difference of black hole masses estimated with these two relations increases with the radio-loudness for extremely radio-loud AGNS, which is consistent with the fact that their equivalent widths of H$_{\\beta}$ emission line become smaller at higher radio-loudness. If the small H$_{\\beta}$ equivalent widths of extremely radio-loud AGNs are indeed caused by the beaming effect, we argue that the optical emission line luminosity may be a better tracer of ionizing luminosity for blazar-type AGNs and the black hole mass derived with the $R-L_{H_\\beta}$ relation are probably more accurate. ", "introduction": "Supermassive black hole is essential for AGNs activities (Lynden-Bell 1969; Rees 1984). The black hole masses of some nearby AGNs have been recently estimated by the reverberation mapping technique (Wandel, Peterson \\& Malkan 1999; Ho 1999; Kaspi \\etal 2000), with which the size of the broad line region (BLR) can be measured from the time delay between the flux variations of the continuum and the emission lines of AGNs. The black hole mass is then estimated using the Virial theorem from the BLR size and the characteristic velocity (determined by the full width at half-maximum (FWHM) of emission line). So far, the reverberation studies have yielded the black hole masses of about 20 Seyfert 1 galaxies and 17 nearby bright quasars. An empirical relation between the BLR size ($R$) and the optical continuum luminosity at 5100$\\AA$ ($L_{5100\\AA}$) has been derived by Kaspi et al. (2000) using the observed data of 34 nearby AGNs. Because the measurement of the BLR size with the reverberation mapping technique needs long-term monitoring of continuum and emission line fluxes, it is impractical for most AGNs. Therefore, the empirical relation has been frequently adopted to estimate the BLR size and then derive the black hole masses for AGNs in some samples of mostly radio-quiet objects (Laor 2000; McLure \\& Dunlop 2001; Wandel 2002), and of purely radio-loud objects (Lacy et al. 2001; Gu, Cao \\& Jiang 2001; Oshlack, Webster \\& Whiting 2002). However, the optical luminosity of some radio-loud AGNs (especially blazars), may not be a good indicator of ionizing luminosity, which is usually related to the UV/optical radiation from the accretion disk around the central black hole. The relativistic jets of blazar-type AGNs not only dominate the radio and high energy X-ray and $\\gamma$ ray radiation, but also significantly contribute to the optical luminosity in some cases (Scarpa \\& Urry 2002). For example, many optical jets have been discovered recently in AGNs by the HST (Scarpa et al. 1999; Jester 2003; Parma \\etal 2003), which clearly suggests that the jets contribute significantly in the optical band. Furthermore, optical synchrotron radiation has been detected for many other radio-loud AGNs (Whiting, Webster \\& Francis 2001; Chiaberge, Capetti, \\& Celloti 2002; Cheung \\etal 2003). Therefore, the measured optical continuum luminosity of some extremely radio-loud AGNs is significantly contributed by the optical radiation from the jets and may be much larger than the ionizing luminosity required to produce broad emission lines. Using the empirical relation between the BLR size and optical luminosity at 5100$\\AA$, which was obtained based on the sample of mostly radio-quiet AGNs (Kaspi \\etal 2000), one would significantly overestimate the actual BLR size and hence the black hole mass of these radio-loud AGNs. Oshlack et al. (2002) have shown that their estimated black hole masses would be lower if the synchrotron contribution to the optical flux is subtracted. However, it is not easy to make such a correction for a large sample of radio-loud AGNs. In addition, the contribution of the host galaxy to the optical continuum should also be taken into account especially when the host galaxy of AGNs can be resolved optically. Therefore, optical luminosity may not be a good indicator of photoionization luminosity of AGNs in some cases. In this paper we will first derive an empirical relation between the BLR size and the H$_\\beta$ emission line luminosity for 34 AGNs in the sample of reverberation mapping studies. We then argue that the BLR size obtained from the H$_\\beta$ luminosity is more reasonable at least for some extremely radio-loud AGNs. Finally we apply this new empirical relation to estimate the black hole masses of some quasars and compare them with previous results. ", "conclusions": "Using the empirical relation between the BLR size and optical continuum luminosity possibly induces an overestimation of the BLR size and hence the black hole mass of some extremely radio-loud AGNs because of the jet contribution to the optical luminosity. We derived another empirical relation between the BLR size and the emission line luminosity, and demonstrated that it can be used to estimate the BLR size and black hole mass of both radio-quiet and radio-loud AGNs. If the relativistic jets and host galaxy have significant contributions to the optical continuum, the emission line luminosity is probably a better tracer of ionizing luminosity. Comparisons of the estimated black hole masses with these two different empirical relations clearly indicate that the difference becomes significant if the radio-loudness of AGNs is larger. Using the $R-L_{H_\\beta}$ relation may result in more accurate estimations of black hole masses of some blazar-type AGNs. In our study we focused on the possible effects of beaming on the optical continuum in radio-loud AGNs and ignored the difference in BLR physics between radio-loud and radio-quiet AGNs. The modest correlation between the EW(H$_{\\beta}$) and the radio-loudness may indicate the presence of beaming effects, though some other effects such as a lower covering factor of the BLR of radio-loud AGNs can also lead to smaller EW (H$_{\\beta}$) values. Because currently we know little about the difference of BLR physics between radio-loud and radio-quiet AGNs, to prove the validity of our approach it is necessary to compare our estimated black hole mass with an independent estimate, for example, from the correlations of black hole mass with central velocity dispersion and host galaxy luminosity. Unfortunately, not many measured values of central velocity dispersion or host galaxy luminosity for extremely radio-loud AGNs are available. Although there are 8 objects in the sample of Brotherton (1996) with measured host magnitude (McLure \\& Dunlop 2001), the radio-loudness of these objects are mostly smaller than 1000 and thus the difference estimated with the $R-L_{H_\\beta}$ and $R-L_{5100\\AA}$ relations is rather small. The velocity dispersion measurements for radio-loud quasars are not available and the [OIII] profile in radio-loud AGNs may not be adopted to estimate the central velocity dispersion because of its complexity. Therefore, further imaging studies on the host galaxy and spectroscopic measurements of the central velocity dispersions of a large sample of extremely radio-loud quasars are still desired to confirm our results. The advantage of using the $R-L_{H_\\beta}$ relation is that we can estimate the black hole mass of AGNs with only two observed parameters, namely the H$_\\beta$ line luminosity and its FWHM, and it can be applied to a larger sample of AGNs with redshift smaller than 0.8. In principle, one can analogously investigate the relation between the BLR size and the luminosity of some ultraviolet emission lines such as MgII and CIV, which may be used to estimate the black hole mass of some high redshift AGNs. Some recent studies have suggested to use the ultraviolet continuum luminosity and the FWHM of ultraviolet emission lines to estimate the black hole mass of high redshift AGNs (Vestergaard 2002; McLure \\& Jarvis 2002). However, the ultraviolet continuum luminosity can similarly suffer the serious contaminations from jet and Blamer continuum, therefore the luminosity of ultraviolet emission line again may be a better indicator of ionizing luminosity than the ultraviolet continuum luminosity. Finally, one should be cautious to the uncertainties in estimating the black hole mass of AGNs using the $R-L_{H_\\beta}$ relation. Firstly, the variations of H$_\\beta$ emission line flux and its FWHM are common in AGNs. Estimating the black hole mass with the values of these two parameters in a single spectrum may lead to large errors. Secondly, the different inclination of the BLR may also significantly affect the results (McLure \\& Dunlop 2001; Wu \\& Han 2001). If the BLR has a flatten geometry and the inclination of BLR is rather small, our derived values of black hole mass may significantly underestimated. However, the ratio of black hole masses estimated with two empirical relations does not depend on the inclination. Better understandings of BLR geometry and dynamics are absolutely needed to diminish the uncertainties in deriving the black hole mass of AGNs (Krolik 2001)." }, "0403/astro-ph0403133_arXiv.txt": { "abstract": "To understand the nuclear stellar populations and star formation histories of the nuclei of spiral galaxies, we have obtained K-band nuclear spectra for 41 galaxies and H-band spectra for 20 galaxies in the ISO Atlas of Bright Spiral Galaxies. In the vast majority of the subsample (80\\%), the near-infrared spectra suggest that evolved red stars completely dominate the nuclear stellar populations and that hot young stars are virtually non-existent. The signatures of recent star formation activity are only found in 20\\% of the subsample, even though older red stars still dominate the stellar populations in these galaxies. Given the dominance of evolved stars in most galaxy nuclei and the nature of the emission lines in the galaxies where they were detected, we suggest that nuclear star formation proceeds in the form of instantaneous bursts. The stars produced by these bursts comprise only $\\sim$2 \\% of the total nuclear stellar mass in these galaxies, but we demonstrate how the nuclear stellar populations of normal spiral galaxies can be built up through a series of these bursts. The bursts were detected only in Sbc galaxies and later, and both bars and interactions appeared to be sufficient but not necessary triggers for the nuclear star formation activity. The vast majority of galaxies with nuclear star formation were classified as HII galaxies. With one exception, LINERs and transition objects were dominated by older red stars, which suggested that star formation was not responsible for generating these galaxies' optical line emission. ", "introduction": "} The nuclear stellar populations of galaxies can reveal important information on the history of nuclear star formation, which could then reveal the mechanisms behind triggering star formation. Understanding nuclear stellar populations is therefore important to understanding the evolution of galaxies and the enhancement of star formation in the universe. Near-infrared spectroscopy is ideal for studying stellar populations because it traces not only the presence of hot young stars though near-infrared hydrogen recombination lines, such as the Brackett $\\gamma$ line, but also the presence of red stars, particularly red supergiants, and the presence of shocks from supernovae activity. Red supergiants and giants can be determined from the presence of various absorption features in the H- and K-bands, such as the CO and Si~{\\small I} absorption features in the H-band and the CO bands longward of 2.29~$\\mu$m \\citep{omo93}. Supernova activity can be inferred from the presence of Fe~{\\small II} emission lines associated with the supernova activity \\citep{omd89, getal91, c93, fw93, getal97, mds02, aetal03} and possibly even molecular hydrogen line emission \\citep{omd89, getal91, fw93} (although molecular hydrogen line emission may be produced by other sources; see \\citet{m94} and \\citet{vr97}). These diagnostics can be used together to characterize the different ages of the various stellar populations within galactic nuclei. Once the nuclear stellar populations of spiral galaxies are well defined, they can be used to answer a number of specific scientific questions. To begin with, this will illuminate how star formation proceeds in the nuclei of ``normal'' spiral galaxies, and whether that star formation is either continuous or a series of short bursts. Once we understand the functionality of star formation over time, we can determine how the nuclei of spiral galaxies have evolved. We can also determine how morphological features, particularly bars, or how environmental influences, such as interactions, influence nuclear star formation activity. Additionally, we can probe the link between star formation activity and Seyfert or LINER activity. Much of the work at near-infrared wavelengths has focues on unusual objects with either AGN or very strong star formation activity. Recent surveys include surveys of ultraluminous infrared galaxies \\citep{getal95, metal99, msmetal01, bwd01}, luminous infrared galaxies \\citep{gjdetal97}, starbursts \\citep{e97, i00, cdd01}, Seyferts \\citep{i00, stl01, rkp02, betal02}, LINERs \\citep{letal98, aetal00, stl01}, and interacting galaxies \\citep{var98}. However, relatively little near-infrared spectroscopic survey has been done with ``normal'' nearby spiral galaxies (see \\citet{mbpetal01} for an example). Such surveys are necessary, though, for understanding nuclear star formation histories, for identifying the triggers that enhance nuclear star formation, and for setting a baseline for the H- and K-band emission from normal spiral galaxies. Without such a baseline, the contribution of relatively quiescent stellar populations to H- and K-band emission in starbursts as well as the older stars' contribution to the overall mass and luminosities in these systems will remain unknown, and the enhancement in star formation activity in exotic systems like ultraluminous infrared galaxies will have no context. Therefore, we have undertaken a near-infrared spectroscopic survey of a subset of the galaxies in the ISO Atlas of Bright Spiral Galaxies (\\cite{betal02a}, henceforth referred to as Paper 1) with the purpose of understanding the nuclear stellar populations and star formation histories of these objects. The data includes K-band spectroscopy for 41 galaxies, with additional H-band spectroscopy for 20 galaxies. We first discuss in Section~\\ref{s_data} the sample, the observations, and the data processing. Next, we divide the galaxies into two groups: quiescent galaxies and non-quiescent galaxies. In Section~\\ref{s_quies}, the spectra of the quiescent galaxies are described in detail, combined together to make a composite quiescent spectrum, and analyzed using population synthesis models. The non-quiescent galaxies are also described in detail in Section~\\ref{s_nonquies}, with particular focus on how their nuclear star formation activity may be related to morphology, environment, and AGN activity. After this, we determine the relative fractions of stars formed from starbursts in these systems. Finally, in Section~\\ref{s_specphotcomp}, we compare quiescent and non-quiescent galaxies in plots of the $\\frac{f_{12\\mu m}}{f_K}$ ratio to the $f_{12\\mu m}$ or $f_K$ luminosities for the inner $15\\arcsec$ as an effort to justify using mid-infrared fluxes normalized by K-band fluxes (used in \\citet{betal02b}, henceforth referred to as Paper 2) to trace star formation activity. ", "conclusions": "\\subsection{Summary of Results} In this sample, 33 of the 41 galaxies nuclei were found to have quiescent spectra. These spectra consisted of continuum emission with many metal and CO absorption lines, which are characteristic of evolved red stars. Using Starburst99, we found that the composite spectrum made from these quiescent galaxies' spectra was best represented by an instantaneous burst that took place $\\sim$180~Myr ago, although we caution that this is not necessarily supposed to reflect the true star formation history. We found few differences between using Salpeter and other IMFs, but we did find that the continuous star formation scenario failed to reproduce the observed equivalent widths. The remaining 8 galaxies have nuclear spectra that include emission lines. In many cases, the underlying continuum exhibited the same slope and contained the same absorption features that were present in the composite quiescent spectra. On top of this quiescent continuum, the galaxy may produce either photoionization lines such as Brackett-$\\gamma$ and He~{\\small I} emission, shock excitation lines such as H$_2$ lines and Fe~{\\small II} emission, or both. The lines present would reflect the age of the young stellar population, with photoionization lines present up to 8~Myr after a burst and shock excitation lines present from 3.5 to 36~Myr after a burst. Some of the systems showed relatively point-like spectral line emission, whereas others exhibited emission in complex structures. One galaxy, where the nuclear emission was quiescent, had a strong extranuclear starburst that was detected and included in some parts of the analysis. These non-quiescent galaxies share many characteristics. All of them are Sbc galaxies or later, which suggests that these instantaneous nuclear bursts are more likely in the later-type galaxies. 6 of the 8 galaxies were definitely barred, which suggested that bars may play a role in triggering nuclear star formation. However, 6 of the 8 galaxies either were interacting or had been involved in interactions, which suggests that interactions are just as important as bars in triggering nuclear star formation. In relation to AGN activity, it appears that most non-quiescent galaxies are H~{\\small II} galaxies and that most LINERs and transition objects are quiescent. The average ratio of young to old stellar masses for the non-quiescent systems was found to be $\\sim$2~\\%. Given that these 2~\\% enhancements occur in $\\sim$20~\\% of the galaxies in the sample, these enhancements are found to be consistent with nuclear stellar populations forming through a series of instantaneous bursts. Finally, we found that the $\\frac{f_{MIR}}{f_K}$ ratios functioned very well in separating non-quiescent galaxies from quiescent galaxies. The exception was with galaxies that contained only shock excitation lines, but the ratio was expected to decrease signficantly for these galaxies. The data and results presented here now form a baseline for comparison to other galaxies with more exotic star formation or AGN activity. Furthermore, the composite quiescent spectra and the conversion factors developed here should provide additional analysis tools for studying enhanced star formation activity. \\subsection{Future Work} While the spectroscopy presented here surveys a range of different kinds of galaxies, it is only a limited sample. A broad range of morphologies were included in the sample, but not a broad range of AGN types; the spectroscopic sample only contained two Seyferts. Furthermore, as Virgo Cluster galaxies were left out of the original sample, no data are available for comparing field and cluster galaxies. A larger near-infrared spectroscopic survey of a complete, volume-limited sample of galaxies would provide additional constraints on these stochastic bursts of star formation, such as the frequency of their occurence, their relation to bars, environment, Seyfert activity, and LINER activity, and the range of masses formed within these bursts. Single slit spectrometers are good for sampling the pointlike nuclei of many nearby galaxies and can return spectroscopic data that accurately represent the stellar populations and total star formation activity within these galaxies. However, when the nuclear star formation structures become more complex, as is the case for NGC~3556 and NGC~4100, then a single slit could possibly miss much of the spectral line emission coming from within the centers of these galaxies. Therefore, integral field near-infrared spectroscopic surveys of nearby spiral galaxies would appear to be warranted. Such a survey would not only provide more accurate data on the stellar populations and total line emission from the nuclei of galaxies but also provide data on the number of galaxies with complex, extended nuclear structures that could lead to additional clues on the phenomenology of nuclear star formation activity." }, "0403/astro-ph0403419_arXiv.txt": { "abstract": "The recently discovered $z=10$ galaxy (Pello \\etal\\ 2004) has a strong Ly$\\alpha$ emission line that is consistent with being surprisingly symmetric, even given the relatively poor quality of its spectrum. The blue wing of a Ly$\\alpha$ line originating at high redshift should be strongly suppressed by resonant hydrogen absorption along the line of sight, an expectation borne out by the observed asymmetric shapes of the existing sample of \\lya\\ emitting sources at lower redshifts ($3< z < 6.7$). Absorption on the blue side of the line of the Pello \\etal\\ source could be reduced if the intergalactic medium (IGM) in the vicinity of the galaxy is highly ionized, but we show that this requires an unrealistically high ionizing emissivity. We suggest instead that the Ly$\\alpha$ emitting gas be receding relative to the surrounding gas with a velocity of $\\gsim 35$ km/s, a large velocity that is plausible only if the galaxy is part of a larger system (group of galaxies) with a velocity dispersion $\\gsim 35$ km/s. We find that with this velocity shift, the observed strength and shape of the line is still consistent with the galaxy being surrounded by its own Str\\\"omgren sphere embedded in a fully neutral IGM. More generally, we predict that at any given redshift, the bright Ly$\\alpha$ emitters with broader lines would exhibit stronger asymmetry than fainter ones. Bright galaxies with symmetric Ly$\\alpha$ lines may be signposts for groups and clusters of galaxies, within which they can acquire random velocities comparable to or larger than their linewidths. ", "introduction": "\\label{sec:introduction} Three independent observations combine to paint a complex picture of the cosmological reionization process. First, the recent quasar absorption spectrum observations by the Sloan Digital Sky Survey show strong evidence that the reionization process completes at $z\\sim 6$ (Becker \\etal\\ 2001; Fan \\etal\\ 2002; Cen \\& McDonald 2002). Second, the latest {\\it Wilkinson Microwave Anisotropy Probe (WMAP)} observations detect a high Thomson scattering optical depth, suggesting that the intergalactic medium (IGM) experienced a significantly ionized state at high redshift somewhere between $z=15-25$ (Kogut \\etal\\ 2003) for (at least) a significant redshift interval. This is somewhat contradicted by the third observational line of evidence of the intergalactic medium having a relatively high temperature at $z\\sim 3-4$, which requires a reionization epoch no earlier than redshift $z=9-10$ (Hui \\& Haiman 2003; Theuns \\etal\\ 2002). While the overall picture is consistent with a pre-WMAP, physically motivated double reionization model (Cen 2003; Wyithe \\& Loeb 2003), a detailed probe of the ionization state of the IGM at high redshift is sorely wanted. Ly$\\alpha$ emission lines from high--redshift sources can serve as probes of the ionization state of the IGM. The damping wing of the Gunn--Peterson (GP) absorption from the IGM can cause a characteristic absorption feature (Miralda-Escud\\'e 1998). For a Ly$\\alpha$ emitting galaxy embedded in a partly neutral IGM, the absorption produces conspicuous effects, i.e. attenuating the emission line, making it asymmetric, and shifting its apparent peak to longer wavelengths (Haiman 2002; Santos 2003). In practice, the expectation is that strong conclusions cannot be drawn from a single galaxy. For example, the relatively strong Ly$\\alpha$ line of the $z=6.6$ galaxy discovered by Hu \\etal (2002) is still consistent with being embedded in a neutral IGM, but surrounded by its own ionized Str\\\"omgren sphere (Haiman 2002; Santos 2003). The recent claim of the detection of a Ly$\\alpha$ emitting galaxy at $z=10$ (Pello \\etal\\ 2004; hereafter P04) provides a new opportunity to study the IGM at $z=10$. This source is especially interesting, since at its high inferred redshift, absorption by the IGM should increase significantly. In this {\\it Letter}, we examine both the observed shape and overall attenuation of the detected Ly$\\alpha$ line, in models where the line is processed through the IGM. We find that in order to achieve the observed symmetry of the Ly$\\alpha$ line profile, the emitting gas in the galaxy must be receding faster than the surrounding gas by at least $35$km/s. Given this required recessional velocity, we find that the P04 source is marginally consistent with being embedded in a fully neutral IGM at $z=10$, with the line suffering an attenuation by a factor of 30. Throughout this paper, we assume the background cosmology to be flat $\\Lambda$CDM with $(\\Omega_\\Lambda,\\Omega_{\\rm m},\\Omega_{\\rm b}, h)=(0.7,0.3,0.04,0.7)$. ", "conclusions": "We considered the implications of the detection of the symmetric, highly attenuated Ly$\\alpha$ emission line from a candidate $z=10$ galaxy. We find that the observed symmetry of the Ly$\\alpha$ emission line can be accounted for if the emitting galaxy is receding relative to the surrounding absorbing gas by a velocity of at least $35$km/s. Such a relative velocity is mostly plausibly achieved if this detected galaxy is a member of a larger system with a velocity dispersion in excess of $35$km/s. Thus, while the difficulties and challenges associated with such observations are formidable, it is not a great surprise, in principle, to be able to detect galaxies with such Ly$\\alpha$ emission lines at high redshift. However, with the required recessional velocity, this galaxy does not place a strong constraint on the ionization state of the IGM, given various uncertainties in the current data and lack of handle of the intrinsic absorption. A fully neutral universe, while not preferred, is still consistent with the observation. A moderate increase in the sample size of such high redshift galaxies will be highly valuable in statistical inferences for the ionization state of the IGM, based on the systematic dependence of the line properties on redshift and luminosity (Haiman 2002; Rhoads \\& Malhotra 2002). More urgent in the short term is to obtain a higher quality spectrum to better characterize the line profile." }, "0403/astro-ph0403305_arXiv.txt": { "abstract": "\\noindent The Generalized Chaplygin Gas (GCG) with the equation of state $p = - \\frac{A}{{\\rho}^{\\alpha}}$ was recently proposed as a candidate for dark energy in the Universe. In this paper we confront the GCG with SNIa data using avaliable samples. Specifically we have tested the GCG cosmology in three different classes of models with (1) $\\Omega_m= 0.3$, $\\Omega_{Ch}= 0.7$; (2) $\\Omega_m= 0.05$, $\\Omega_{Ch}= 0.95$ and (3) $\\Omega_m = 0$, $\\Omega_{Ch} = 1$, as well as a model without prior assumptions on $\\Omega_m$. The best fitted models are obtained by minimalizing the $\\chi^2$ function. We supplement our analysis with confidence intervals in the $(A_0, \\alpha)$ plane by marginalizing the probability density functions over remaining parameters assuming uniform priors. We have also derived one-dimensional probability distribution functions for $\\Omega_{Ch}$ obtained from joint marginalization over $\\alpha$ and $A_0$. The maximum value of such PDF provides the most probable value of $\\Omega_{Ch}$ within the full class of GCG models. The general conclusion is that SNIa data give support to the Chaplygin gas (with $\\alpha = 1$). However noticeable preference of $A_0$ values close to 1 means that the $\\alpha$ dependence becomes insignificant. It is reflected on one dimensional PDFs for $\\alpha$ which turned out to be flat meaning that the power of present supernovae data to discriminate between various GCG models (differing by $\\alpha$) is weak. Extending our analysis by relaxing the prior assumption of the flatness of the Universe leads to the result that even though the best fitted values of $\\Omega_k$ are formally non-zero, still they are close to the flat case. Our results show clearly that in GCG cosmology distant (i.e. $z >1$) supernovae should be brighter than in $\\Lambda$CDM model. Therefore one can expect that future supernova experiments (e.g., SNAP) having access to higher redshifts will eventually resolve the issue whether the dark energy content of the Universe could be described as a the Chaplygin gas. Moreover, it would be possible to differentiate between models with various value of $\\alpha$ parameter and/or discriminated between GCG, Cardassian and $\\Lambda$CDM models. This discriminative power of the forthcoming mission has been demonstrated on simulated SNAP data. ", "introduction": "For a couple of years two independent observational programs --- the high redshift supernovae surveys (Perlmutter et al. 1999) and CMBR small scale anisotropy measurements (de Bernardis et al. 2000, Benoit et al. 2003, Hinshaw et al. 2003) have brought a new picture of the Universe in the large. While interpreted ithin the FRW models results of these programs suggest that our Universe is flat (as inferred from the location of acoustic peaks in CMBR power spectrum) and presently accelerates its expansion (as inferred from the SNIa Hubble diagram). Combined with the independent knowledge about the amount of baryons and CDM estimated to be $\\Omega_m = 0.3$ (Turner 2002) it follows that about $\\Omega_X = 0.7$ fraction of critical density $\\rho_{cr} = \\frac{3 c^2 H_0^2}{8 \\pi G}$ should be contained in a mysterious component called ``dark energy''. The most obvious candidate for this smooth component permeating the Universe is the cosmological constant $\\Lambda$ representing the energy of the vacuum. Well known fine tuning problems led many people to seek beyond the $\\Lambda$ framework, and the concept of the quintessence had been conceived. Usually the quintessence is described in a phenomenological manner, as a scalar field with an appropriate potential (Ratra \\& Peebles 1988, Caldwell, Dave \\& Steinhardt 1995, Frieman, Stebbins \\& Waga 1995). It turns out, however, that quintessence program also suffers from its own fine tuning problems (Kolda \\& Lyth 1999). In 1904 Russian physicist Chaplygin introduced the exotic equation of state $p = - \\frac{A}{\\rho}$ to discribe an adiabatic aerodynamic process (Chaplygin 2004). The attractiveness of this equation of state in the context dark energy models comes mainly from the fact that it gives the unification of both dark energy (postulated in cosmology to explain current aceleration of the Universe) and clustered dark matter which is postulated in astrophysics to explain the flat rotation curves of spiral galaxies. It is interesting that the Chaplygin gas can be derived from the quintessence Lagrangian for the scalar field $\\phi$ with some potential and also from the Born-Infeld form of the Lagrangian (Kamenshchik, Moschella \\& Pasquier 2001). The Chaplygin equation of state has some interesting connections with string theory and it admits the interpretation in the framework of brane cosmologies (Jackiw 2000). Recently the so called Chaplygin gas (Kamenshchik, Moschella \\& Pasquier 2001, Fabris, Gon{\\c c}alves \\& de Souza 2002, Szyd{\\l}owski \\& Czaja 2004) --- a hypothetical component with the equation of state $p = - \\frac{A}{\\rho}$ --- was proposed as a challenge to the above mentioned candidates for dark energy. This, also purely phenomenological, entity has interesting connections with string theory (Ogawa 2000). Currently its generalizations admitting the equation of state $p= - \\frac{A}{\\rho^{\\alpha}}$ where $0\\le \\alpha \\le 1$ have been proposed (Bento, Bertolami \\& Sen 2002, Carturan \\& Finelli 2002a). In this paper we confront the Generalized Chaplygin Gas with the SNIa data. At this point our choice of Generalized Chaplygin Gas cosmologies deserves a sort of justification. There are two approaches in the literature. First one is phenomenological, namely having no preferred theory of dark energy responsible for acceleration of the Universe one characterizes it as a cosmic fluid with an equation of state $p_X = w \\rho_X$ where $w \\geq -1$ (see e.g. (Chiba 1998,Turner \\& White 1997) and an immense literature that appeared thereafter). Because, as already mentioned above, a strain of ideas about dark energy is associated with an evolving scalar there are good reasons to expect that cosmic equation of state could be time dependent i.e. $ w = w(t) = w(z)$ (e.g. Weller \\& Albrecht 2001, Maor et al. 2001 and many others thereafter). This approach seems attractive from the perspective of analyzing observational data such like supernovae surveys and indeed this approach was taken while first analyzing the data (Perlmutter et al. 1999, Knop et al. 2003 or Riess et al. 2004). However even though such analysis places constrains on {\\it any} potential theory that might explain the dark energy phenomenon, ultimately one always ends up at testing a {\\it specific} theory. Along this line there appeared attempts to reconstruct the scalar potential, if the scalar field was responsible for dark energy (e.g. Alam et al. 2003 and references therein). Our approach goes along this philosophy but instead is devoted to the Generalized Chaplygin Gas which is being recently considered as candidate to unified dark matter-energy component (i.e. responsible for both clustering and accelerated expansion (Makler, de Oliveira \\& Waga 2003). The cosmological models with the Generalized Chaplygin Gas have also many special features which make them atractive. In standard cosmological model one can clearly distinguish the epochs of radiation domination followed by (ordinary) matter domination (with decelerated expansion). As mentioned above supernova data suggest that the epoch of decelerated expansion ended and switched to accelerated epoch --- dominated by dark energy. The Generalized Chaplygin Gas models describe smoothly the transition from the decelerated to accelerated epochs. They represent the simplest deformation of concordance $\\Lambda$CDM (Gorini et al. 2004). And moreover, they propose a new unified macroscopic (phenomenological) description of both dark energy and dark matter. This places them in a distinguished position from the point of view of Occam's razor principle. It should be also noted that the Genaralized Chaplygin Gas model allowed us to explain presently observed acceleration of the Universe without the cosmological constant and/or modification of Einstein's equations. If one takes seriously given dark energy scenario (necessary to explain cosmic acceleration) one should also consider the behaviour of perturbations in such a universe. In the framework of quintessence models with the barotropic equation of state (i.e. $p= w \\rho$ and $w=const$) one faces the problem of instabilities in short scales. This appears because the speed of sound squared (equal here to $w$) is negative (and constant). Calculation of the sound speed in Generalized Chaplygin Gas model (see below) reveal its non-barotropic nature. The perturbations in GCG models are stable in short scales even in an accelerating phase (Carturan \\& Finelli 2002a). Moreover, they behave like dust perturbations when Chaplygin Gas is in dust regime. Another motivation for studying Generalized Chaplygin Gas models goes from theoretical physics --- specifically from attempts to describe the dark energy in terms of the Lagrangian for a tachyonic field (Garousi 2000, Sen 2002). Of course it would be nice to have a description of dark energy in terms of the non quintessence Lagrangian as it describes the nature of dark energy while the cosmological constant is only phenomelogical and effective description. One should also note that the Generalized Chaplygin Gas equation of state arises in modern physics in the context of brane models (Bordemann \\& Hoppe 1993, Kamenshchik, Moschella \\& Pasquier 2001, Randall \\& Sundrum 1999) where the Generalized Chaplygin Gas manifests itself as an effect of immersion of our Universe in multidimensional bulk space. Generalized Chaplygin gas models have been intensively studied in the literature and in particular they have been tested against supernovae data (Makler, de Oliveira \\& Waga 2003, Avelino et al. 2003, Collistete et al. 2003), lensing statistics (Dev, Alcaniz \\& Jain 2003), CMBR measurements (Bento, Bertolami \\& Sen 2003a, 20003b, Carturan \\& Finelli 2003b, Amendola et al. 2003), age-redshift relation (Alcaniz, Jain \\& Dev 2003), x-ray luminosities of galaxy clusters (Cunha, Lima \\& Alcaniz 2003) or from the large scale structure considerations (Bean \\& Dor{\\'e} 2003, Multamaki, Manera \\& Gaztanaga 2003, Bilic et al. 2003). Perspectives to distinguish between Generalized Chaplygin Gas, brane-world scenarios and quintessence in forthcoming gravity wave experiments has been discussed in (Biesiada 2003). Although the results are in general mutually consistent there was no strong convergence to unique values of $A_0$, $\\alpha$ parameters characterizing Chaplygin gas equation of state. Makler, de Oliveira \\& Waga (2003) have considered the FRW model filled completely with Generalized Chaplygin Gas and concluded that whole class of such models is consistent with current SNIa data although the value of $\\alpha = 0.4$ is favoured. This result has been confirmed by our analysis (class (3) models). However, when the existing knowledge about baryonic matter content of the Universe was incorporated into the study our results were different from Makler, de Oliveira \\& Waga (2003) who found that $\\alpha = 0.15$ was preferred (assuming $\\Omega_m = 0.04$ which is very close to our assumption for class (2) models). As noticed by Bean \\& Dor{\\'e} (2003) Generalized Chaplygin Gas models have an inherent degeneracy with cosmological constant models as far as background evolution is concerned, and therefore they have a good fit with SNIa data. These degeneracies disappear at the level of evolution of perturbations and hence confrontation with CMBR spectrum would be decisive. Using available data on the position of CMBR peaks measured by BOOMERANG (de Bernardis et al. 2000) and Archeops (Benoit et al. 2003, Hinshaw et al. 2003, Bento, Bertolami \\& Sen (2002) obtained the following constraints: $0.81 \\leq A_0 \\leq 0.85$ and $0.2 \\leq \\alpha \\leq 0.6$ at $68 \\%$ CL in the model representative of our class (2) (i.e. with $\\Omega_m = 0.05$ assumed). Another estimation of the parameter $\\alpha$ was done by Amendola et al. (2003) with WMAP Data. The obtained the $0\\le \\alpha < 0.2$ at $95\\%$ confidence level. Using the angular size statistics for extragalactic sources combined with SNIa data it was found in (Alcaniz \\& Lima 2003) that in the the $\\Omega_m =0.3$ and $\\Omega_{Ch}=0.7$ scenario best fitted values of model parameters are $A_0=0.83$ and $\\alpha=1.$ respectively. Recent paper by Bertolami et al. (2004) in which Generalized Chaplygin Gas models have been analyzed against Tonry et al. (2003) supernovae data relaxing the prior assumption on flatness suggests, surprisingly as the authors admit, the preference of $\\alpha > 1$. \\begin{deluxetable}{@{}p{1.5cm}rrrrrrr} \\tabletypesize{\\scriptsize} \\tablewidth{0pt} \\tablecaption{Results of statistical analysis of Generalized Chaplygin Gas model (with marginalization over ${\\cal M}$) performed on analyzed samples of SNIA (A, C, K6, K3, TBI, TBII, Silver, Gold) as a minimum $\\chi^2$ best-fit (denoted BF) and with the maximum likelihood method (denoted L). First two rows for each sample refer to no prior on $\\Omega_m$. The same analysis was repeated with fixed priors $\\Omega_m=0.0$, $\\Omega_m=0.05$ and $\\Omega_m=0.3$.} \\label{tab:1} \\tablehead{\\colhead{sample} & \\colhead{$\\Omega_m$} & \\colhead{$\\Omega_{Ch}$} & \\colhead{$A_0$} & \\colhead{$\\alpha$} & \\colhead{$\\mathcal{M}$} & \\colhead{$\\chi^2$}& \\colhead{method}} \\startdata A & 0.00 & 1.00 & 0.77 & 1.00 &-3.39 & 95.4 & BF \\\\ & 0.17 & 0.83 & 0.83 & 0.00 &-3.36 & --- & L \\\\ & 0.00 & 1.00 & 0.77 & 1.00 &-3.39 & 95.4 & BF \\\\ & 0.00 & 1.00 & 0.73 & 1.00 &-3.38 & --- & L \\\\ & 0.05 & 0.95 & 0.80 & 1.00 &-3.39 & 95.4 & BF \\\\ & 0.05 & 0.95 & 0.76 & 1.00 &-3.38 & --- & L \\\\ & 0.30 & 0.70 & 0.96 & 1.00 &-3.39 & 95.8 & BF \\\\ & 0.30 & 0.70 & 0.96 & 0.00 &-3.38 & --- & L \\\\ \\tableline % C & 0.00 & 1.00 & 0.80 & 1.00 &-3.44 & 52.9 & BF \\\\ & 0.15 & 0.85 & 0.86 & 0.00 &-3.41 & --- & L \\\\ & 0.00 & 1.00 & 0.80 & 1.00 &-3.44 & 52.9 & BF \\\\ & 0.00 & 1.00 & 0.76 & 0.49 &-3.43 & --- & L \\\\ & 0.05 & 0.95 & 0.83 & 1.00 &-3.44 & 53.0 & BF \\\\ & 0.05 & 0.95 & 0.79 & 0.11 &-3.43 & --- & L \\\\ & 0.30 & 0.70 & 0.99 & 1.00 &-3.42 & 53.3 & BF \\\\ & 0.30 & 0.70 & 0.99 & 0.00 &-3.39 & --- & L \\\\ \\tableline K6 & 0.00 & 1.00 & 0.81 & 1.00 &-3.52 & 55.3 & BF \\\\ & 0.10 & 0.90 & 0.88 & 0.00 &-3.51 & --- & L \\\\ & 0.00 & 1.00 & 0.81 & 1.00 &-3.52 & 55.3 & BF \\\\ & 0.00 & 1.00 & 0.78 & 0.71 &-3.52 & --- & L \\\\ & 0.05 & 0.95 & 0.84 & 1.00 &-3.52 & 55.4 & BF \\\\ & 0.05 & 0.95 & 0.81 & 0.06 &-3.52 & --- & L \\\\ & 0.30 & 0.70 & 1.00 & 1.00 &-3.51 & 55.9 & BF \\\\ & 0.30 & 0.70 & 1.00 & 0.00 &-3.49 & --- & L \\\\ \\tableline K3 & 0.00 & 1.00 & 0.85 & 1.00 &-3.48 & 60.4 & BF \\\\ & 0.11 & 0.89 & 0.88 & 0.00 &-3.45 & --- & L \\\\ & 0.00 & 1.00 & 0.85 & 1.00 &-3.48 & 60.4 & BF \\\\ & 0.00 & 1.00 & 0.80 & 0.30 &-3.47 & --- & L \\\\ & 0.05 & 0.95 & 0.87 & 1.00 &-3.47 & 60.4 & BF \\\\ & 0.05 & 0.95 & 0.84 & 0.00 &-3.47 & --- & L \\\\ & 0.30 & 0.70 & 1.00 & 1.00 &-3.44 & 61.4 & BF \\\\ & 0.30 & 0.70 & 1.00 & 0.00 &-3.42 & --- & L \\\\ \\tableline TBI & 0.00 & 1.00 & 0.79 & 1.00 &15.895&273.9 & BF \\\\ & 0.00 & 1.00 & 0.81 & 1.00 &15.905& --- & L \\\\ & 0.00 & 1.00 & 0.79 & 1.00 &15.895&273.8 & BF \\\\ & 0.00 & 1.00 & 0.75 & 1.00 &15.905& --- & L \\\\ & 0.05 & 0.95 & 0.82 & 1.00 &15.895&274.0 & BF \\\\ & 0.05 & 0.95 & 0.78 & 1.00 &15.915& --- & L \\\\ & 0.30 & 0.70 & 0.97 & 1.00 &15.915&275.8 & BF \\\\ & 0.30 & 0.70 & 0.96 & 0.00 &15.915& --- & L \\\\ \\tableline TBII & 0.00 & 1.00 & 0.78 & 1.00 &15.915&186.5 & BF \\\\ & 0.00 & 1.00 & 0.81 & 1.00 &15.925& --- & L \\\\ & 0.00 & 1.00 & 0.78 & 1.00 &15.915&186.5 & BF \\\\ & 0.00 & 1.00 & 0.75 & 1.00 &15.915& --- & L \\\\ & 0.05 & 0.95 & 0.81 & 1.00 &15.915&186.6 & BF \\\\ & 0.05 & 0.95 & 0.78 & 1.00 &15.925& --- & L \\\\ & 0.30 & 0.70 & 0.97 & 1.00 &15.925&188.4 & BF \\\\ & 0.30 & 0.70 & 0.96 & 0.00 &15.935& --- & L \\\\ \\tableline Silver& 0.00 & 1.00 & 0.82 & 1.00 &15.945&229.4 & BF \\\\ & 0.00 & 1.00 & 0.84 & 1.00 &15.945& --- & L \\\\ & 0.00 & 1.00 & 0.82 & 1.00 &15.945&229.4 & BF \\\\ & 0.00 & 1.00 & 0.79 & 1.00 &15.955& --- & L \\\\ & 0.05 & 0.95 & 0.85 & 1.00 &15.945&229.6 & BF \\\\ & 0.05 & 0.95 & 0.81 & 1.00 &15.955& --- & L \\\\ & 0.30 & 0.70 & 0.99 & 1.00 &15.965&232.3 & BF \\\\ & 0.30 & 0.70 & 0.99 & 0.00 &15.965& --- & L \\\\ \\tableline Gold & 0.00 & 1.00 & 0.81 & 1.00 &15.945&173.7 & BF \\\\ & 0.00 & 1.00 & 0.83 & 1.00 &15.955& --- & L \\\\ & 0.00 & 1.00 & 0.81 & 1.00 &15.945&173.7 & BF \\\\ & 0.00 & 1.00 & 0.77 & 1.00 &15.955& --- & L \\\\ & 0.05 & 0.95 & 0.84 & 1.00 &15.945&173.8 & BF \\\\ & 0.05 & 0.95 & 0.80 & 1.00 &15.955& --- & L \\\\ & 0.30 & 0.70 & 0.99 & 1.00 &15.965&175.6 & BF \\\\ & 0.30 & 0.70 & 0.99 & 0.00 &15.965& --- & L \\\\ \\enddata \\end{deluxetable} \\begin{table} \\caption{Generalized Chaplygin Gas model parameter values obtained from the marginal probability density functions calculated on Perlmutter, Knop, Tonry/Barris and Riess samples with $\\Omega_m$ prior relaxed.} \\label{tab:2} \\begin{tabular}{@{}p{1.5cm}rrrr} \\hline \\hline sample & $\\Omega_m$ & $\\Omega_{Ch}$ & $A_0$ & $\\alpha$ \\\\ \\hline A & $ 0.17^{+0.08}_{-0.17}$ & $0.83^{+0.17}_{-0.08}$ & $0.83^{+0.14}_{-0.09}$ & $-0.0^{+0.67}$ \\\\ C & $0.15^{+0.08}_{-0.15}$ & $0.85^{+0.15}_{-0.08}$ & $0.86^{+0.13}_{-0.10}$ & $0.0^{+0.66}$ \\\\ K6& $ 0.10^{+0.11}_{-0.10}$ & $0.90^{+0.10}_{-0.11}$ & $0.88^{+0.12}_{-0.08}$ & $-0.0^{+0.66}$ \\\\ K3& $0.11^{+0.07}_{-0.11}$ & $0.89^{+0.11}_{-0.07}$ & $0.88^{+0.11}_{-0.05}$ & $0.0^{+0.66}$\\\\ TBI& $0.00^{+0.21}$ & $1.00_{-0.21}$ & $0.81^{+0.12}_{-0.07}$ & $1.0_{-0.60}$\\\\ TBII&$0.00^{+0.21}$ & $1.00_{-0.21}$ & $0.81^{+0.12}_{-0.07}$ & $1.0_{-0.62}$\\\\ Silver&$0.00^{+0.18}$ & $1.00_{-0.18}$ & $0.84^{+0.09}_{-0.06}$ & $1.0_{-0.59}$\\\\ Gold& $0.00^{+0.20}$ & $1.00_{-0.20}$ & $0.83^{+0.11}_{-0.07}$ & $1.0_{-0.64}$\\\\ \\hline \\end{tabular} \\end{table} \\begin{table} \\caption{Generalized Chaplygin Gas model parameter values obtained from the marginal probability density functions calculated on Perlmutter, Knop, Tonry/Barris and Riess samples. The analysis was done with fixed $\\Omega_m=0.0$, $\\Omega_m=0.05$ and $\\Omega_m=0.3$.} \\label{tab:3} \\begin{tabular}{@{}p{1.5cm}rrrr} \\hline \\hline sample & $\\Omega_m$ & $\\Omega_{Ch}$ & $A_0$ & $\\alpha$ \\\\ \\hline A &$ 0.00$ & $1.00$ & $0.73^{+0.08}_{-0.10}$ & $1.0_{-0.63}$ \\\\ &$ 0.05$ & $0.95$ & $0.76^{+0.08}_{-0.09}$ & $1.0_{-0.66}$ \\\\ &$ 0.30$ & $0.70$ & $0.96^{+0.04}_{-0.09}$ & $0.0^{+0.65}$ \\\\ C &$ 0.00$ & $1.00$ & $0.76^{+0.08}_{-0.10}$ & $0.49^{+0.36}_{-0.35}$ \\\\ &$ 0.05$ & $0.95$ & $0.79^{+0.08}_{-0.11}$ & $0.41^{+0.27}_{-0.41}$ \\\\ &$ 0.30$ & $0.70$ & $0.99^{+0.01}_{-0.11}$ & $0.0^{+0.64}$ \\\\ K6 &$ 0.00$ & $1.00$ & $0.78^{+0.07}_{-0.09}$ & $0.71^{+0.29}_{-0.40}$ \\\\ &$ 0.05$ & $0.95$ & $0.81^{+0.08}_{-0.09}$ & $0.06^{+0.61}_{-0.06}$ \\\\ &$ 0.30$ & $0.70$ & $1.00_{-0.10} $ & $0.0^{+0.64}$ \\\\ K3 &$ 0.00$ & $1.00$ & $0.80^{+0.06}_{-0.06}$ & $0.30^{+0.39}_{-0.30}$ \\\\ &$ 0.05$ & $0.95$ & $0.84^{+0.05}_{-0.07}$ & $0.0^{+0.67}$ \\\\ &$ 0.30$ & $0.70$ & $1.00_{-0.06} $ & $0.0^{+0.63}$ \\\\ TBI &$ 0.00$ & $1.00$ & $0.75^{+0.04}_{-0.05}$ & $1.0_{-0.54}$ \\\\ &$ 0.05$ & $0.95$ & $0.78^{+0.04}_{-0.06}$ & $1.0_{-0.55}$ \\\\ &$ 0.30$ & $0.70$ & $0.96^{+0.04}_{-0.04}$ & $0.0^{+0.67}$ \\\\ TBII &$ 0.00$ & $1.00$ & $0.75^{+0.04}_{-0.06}$ & $1.0_{-0.54}$ \\\\ &$ 0.05$ & $0.95$ & $0.78^{+0.04}_{-0.06}$ & $1.0_{-0.54}$ \\\\ &$ 0.30$ & $0.70$ & $0.96^{+0.04}_{-0.04}$ & $0.0^{+0.67}$ \\\\ Silver&$ 0.00$ & $1.00$ & $0.79^{+0.03}_{-0.05}$ & $1.0_{-0.52}$ \\\\ &$ 0.05$ & $0.95$ & $0.81^{+0.04}_{-0.04}$ & $1.0_{-0.54}$ \\\\ &$ 0.30$ & $0.70$ & $0.99^{+0.01}_{-0.03}$ & $0.0^{+0.64}$ \\\\ Gold &$ 0.00$ & $1.00$ & $0.77^{+0.04}_{-0.05}$ & $1.0_{-0.58}$ \\\\ &$ 0.05$ & $0.95$ & $0.80^{+0.04}_{-0.05}$ & $1.0_{-0.59}$ \\\\ &$ 0.30$ & $0.70$ & $0.99^{+0.01}_{-0.04}$ & $0.0^{+0.64}$ \\\\ \\hline \\end{tabular} \\end{table} \\begin{table} \\noindent \\caption{Results of statistical analysis of Generalized Chaplygin Gas models with flat prior relaxed and with marginalization over ${\\cal M}$ performed on Knop Samples K3 as a minimum $\\chi^2$ best-fit(denoted BF) and with the maximum likelihood method (denoted L). First two rows refer to no prior on $\\Omega_m$. The same analysis was repeated with fixed $\\Omega_m=0.0$, $\\Omega_m=0.05$ and $\\Omega_m=0.3$.} \\label{tab:4} \\begin{tabular}{@{}p{1.5cm}rrrrrrrr} \\hline \\hline sample & $\\Omega_k$ & $\\Omega_m$ & $\\Omega_{Ch}$ & $A_0$ & $\\alpha$ & $\\mathcal{M}$ & $\\chi^2$& method \\\\ \\hline K3 & -0.19 & 0.00 & 1.19 & 0.82 & 1.00 &-3.48 & 60.3 & BF \\\\ & -0.60 & 0.00 & 1.26 & 0.89 & 0.00 &-3.46 & --- & L \\\\ & -0.25 & 0.00 & 1.25 & 0.82 & 1.00 &-3.49 & 60.3 & BF \\\\ & 0.10 & 0.00 & 0.90 & 0.76 & 0.00 &-3.46 & --- & L \\\\ & -0.28 & 0.05 & 1.23 & 0.84 & 1.00 &-3.49 & 60.3 & BF \\\\ & 0.05 & 0.05 & 0.90 & 0.78 & 0.00 &-3.47 & --- & L \\\\ & -0.48 & 0.30 & 1.18 & 0.93 & 0.97 &-3.49 & 60.3 & BF \\\\ & -0.35 & 0.30 & 1.05 & 0.88 & 0.00 &-3.47 & --- & L \\\\ Gold& -0.12 & 0.00 & 1.12 & 0.80 & 0.99 &15.945&173.4 & BF \\\\ & -0.32 & 0.00 & 1.06 & 0.82 & 0.00 &15.945& --- & L \\\\ & -0.13 & 0.00 & 1.13 & 0.81 & 1.00 &15.935&173.4 & BF \\\\ & -0.19 & 0.00 & 1.19 & 0.76 & 0.85 &15.945& --- & L \\\\ & -0.17 & 0.05 & 1.12 & 0.83 & 1.00 &15.935&173.4 & BF \\\\ & -0.20 & 0.05 & 1.15 & 0.78 & 0.54 &15.945& --- & L \\\\ & -0.31 & 0.30 & 1.01 & 0.94 & 1.00 &15.955&173.6 & BF \\\\ & -0.30 & 0.30 & 1.00 & 0.91 & 0.00 &15.945& --- & L \\\\ \\hline \\end{tabular} \\end{table} \\begin{table} \\caption{Results of statistical analysis of Generalized Chaplygin Gas models with flat prior relaxed and with marginalization over ${\\cal M}$ performed on Knop Samples K3. Model parameter values are obtained from the marginal probability density functions. First row refer to no prior on $\\Omega_m$. The same analysis was repeated with fixed $\\Omega_m=0.0$, $\\Omega_m=0.05$ and $\\Omega_m=0.3$.} \\label{tab:5} \\begin{tabular}{@{}p{1.5cm}rrrrr} \\hline \\hline sample & $\\Omega_k$ & $\\Omega_m$ & $\\Omega_{Ch}$ & $A_0$ & $\\alpha$ \\\\ \\hline K3 & $-0.60^{+0.38}_{-0.38}$ & $ 0.00^{+0.29}$ & $1.26^{+0.25}_{-0.39}$ & $0.89^{+0.11}_{-0.07}$ & $ 0.0^{+0.64}$ \\\\ & $ 0.10^{+0.37}_{-0.60}$ & $ 0.00$ & $0.90^{+0.59}_{-0.37}$ & $0.76^{+0.10}_{-0.07}$ & $ 0.0^{+0.66}$ \\\\ & $ 0.05^{+0.31}_{-0.58}$ & $ 0.05$ & $0.90^{+0.58}_{-0.31}$ & $0.78^{+0.10}_{-0.06}$ & $ 0.0^{+0.66}$ \\\\ & $-0.35^{+0.17}_{-0.40}$ & $ 0.30$ & $1.05^{+0.41}_{-0.17}$ & $0.88^{+0.09}_{-0.05}$ & $ 0.0^{+0.63}$ \\\\ Gold& $-0.32^{+0.25}_{-0.25}$ & $ 0.00^{+0.28}$ & $1.06^{+0.24}_{-0.22}$ & $0.82^{+0.13}_{-0.05}$ & $ 0.0^{+0.64}$ \\\\ & $-0.19^{+0.29}_{-0.28}$ & $ 0.00$ & $1.19^{+0.28}_{-0.29}$ & $0.76^{+0.03}_{-0.05}$ & $ 0.85^{+0.15}_{-0.52}$ \\\\ & $-0.20^{+0.28}_{-0.29}$ & $ 0.05$ & $1.15^{+0.29}_{-0.28}$ & $0.78^{+0.05}_{-0.06}$ & $ 0.54^{+0.36}_{-0.32}$ \\\\ & $-0.30^{+0.21}_{-0.23}$ & $ 0.30$ & $1.00^{+0.23}_{-0.21}$ & $0.91^{+0.04}_{-0.05}$ & $ 0.00^{+0.60}$ \\\\ \\hline \\end{tabular} \\end{table} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f1.eps} \\caption{Residuals (in mag) between the Einstein-de Sitter model (zero line), the $\\Lambda$CDM model (upper curve) and the best-fitted Generalized Chaplygin Gas model with $\\Omega_m= 0.3, \\Omega_{Ch}= 0.7$ (middle curve), sample K3.} \\label{fig:1} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f2.eps} \\caption{Levels of constant $\\chi^{2}$ on the plane $(A_0,\\alpha)$ for Generalized Chaplygin Gas model with $\\Omega_m= 0.3, \\Omega_{Ch}= 0.7$, sample K3, marginalized over ${\\cal M}$. The figure shows preferred values of $A_0$ and $\\alpha$.} \\label{fig:2} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f3.eps} \\caption{Confidence levels on the plane $(A_0,\\alpha)$ for Generalized Chaplygin Gas model with $\\Omega_m= 0.3, \\Omega_{Ch}= 0.7$, sample K3, marginalized over ${\\cal M}$. The figure shows the ellipses of preferred values of $A_0$ and $\\alpha$.} \\label{fig:3} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f4.eps} \\caption{Residuals (in mag) between the Einstein-de Sitter model (zero line), the flat $\\Lambda$CDM model (upper curve) and the best-fitted Generalized Chaplygin Gas model with $\\Omega_m= 0.05, \\Omega_{Ch}= 0.95$ (middle curve), sample K3.} \\label{fig:4} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f5.eps} \\caption{Levels of constant $\\chi^{2}$ on the plane $(A_0,\\alpha)$ for Generalized Chaplygin Gas model with $\\Omega_m= 0.05, \\Omega_{Ch}= 0.95$, sample K3, marginalized over ${\\cal M}$. The figure shows preferred values of $A_0$ and $\\alpha$.} \\label{fig:5} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f6.eps} \\caption{Confidence levels on the plane $(A_0,\\alpha)$ for Generalized Chaplygin Gas model with $\\Omega_m= 0.05, \\Omega_{Ch}= 0.95$, sample K3, marginalized over ${\\cal M}$. The figure shows the ellipses of preferred values of $A_0$ and $\\alpha$.} \\label{fig:6} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f7.eps} \\caption{Residuals (in mag) between the Einstein-de Sitter model (zero line), the flat $\\Lambda$CDM model (upper curve) and the best-fitted Generalized Chaplygin Gas model with $\\Omega_m= 0 , \\Omega_{Ch}= 1$ (middle curve), sample K3.} \\label{fig:7} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f8.eps} \\caption{Levels of constant $\\chi^{2}$ on the plane $(A_0,\\alpha)$ for Generalized Chaplygin Gas model with $\\Omega_m= 0, \\Omega_{Ch}= 1$, sample K3, marginalized over ${\\cal M}$. The figure shows preferred values of $A_0$ and $\\alpha$.} \\label{fig:8} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f9.eps} \\caption{Confidence levels on the plane $(A_0,\\alpha)$ for Generalized Chaplygin Gas model with $\\Omega_m= 0, \\Omega_{Ch}= 1$, sample K3, marginalized over ${\\cal M}$. The figure shows the ellipses of preferred values of $A_0$ and $\\alpha$.} \\label{fig:9} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f10.eps} \\caption{The density distribution (one dimensional PDF) for $\\Omega_{Ch}$ obtained from sample K3 by marginalization over remaining parameters of the model. We obtain the limit $\\Omega_{Ch} > 0.70$ at the confidence level $95.4 \\%$.} \\label{fig:10} \\end{figure} ", "conclusions": "It is apparent that Generalized Chaplygin Gas models have brighter supernovae at redshifts $z>1$. Indeed one can see on respective figures (Figs 1, 4, 7, 13) that systematic deviation from the baseline Einstein de Sitter model gets larger at higher redshifts. This prediction seems to be independent of analysed sample. We obtained that the estimated value of $A_0$ is close to $0.8$ in all considered models with exepction the model class (1) ($\\Omega_m=0.3$) when $A_0>0.95$. Extending our analysis by relaxing the flat prior lead to the result that even though the best fitted values of $\\Omega_k$ are formally non-zero, yet they are close to the flat case. It should be viewed as an advantage of the GCG model since in similar analysis of $\\Lambda$CDM model in Perlmutter et al. (1999) high negative value of $\\Omega_{k}$ were found to be best fitted to the data and independent inspiration from CMBR and extragalactic astronomy has been invoked to fix the curvature problem. Another advantage of GCG model is that in a natural way we obtained the conclusion that matter (baryonic) component should be small what is in agreement with prediction from BBN (Big Bang Nucleosynthesis). Both estimations of $A_0$, $\\Omega_k$ and $\\Omega_m$ are independent of the sample used in our analysis. Our results suggest that SNIa data support the Chaplygin gas (i.e. $\\alpha = 1$) scenario when the $\\chi^2$ best fitting procedure is used. The minimization procedure performed on Tonry/Barris and Riess data gives also $\\alpha=1$ (only except the model with fixed $\\Omega_m=0.3$). However, the maximum likelihood fitting with Knop et al.'s sample prefers, quite unexpectedly, a small value of $\\alpha$ or even $\\alpha = 0$, i.e. the $\\Lambda$CDM scenario. Please note that small value of $\\alpha$ is in agreement with the results obtained from CMBR (de Bernardis et al. 2000 Benoit et al. 2003, Hinshaw et al. 2003, Bento, Bertolami \\& Sen 2002, Amendola et al. 2003) and with the recent analysis of Zhu (2004) who, using combined data of X-ray gas mass fraction of the galaxy cluster, FR IIB radiogalaxies and combined sample Perlmutter et al. (1998) and Riess et al. (1998, 2001), sugested that $\\alpha$ could be even less than 0. The results are dependent both on the sample chosen and on the prior knowledge of $\\cal M$ in which the Hubble constant and intrinsic luminosity of SNIa are entangled. Moreover the observed preference of $A_0$ values close to 1 means that the $\\alpha$ dependence becomes insignificant (see equation (\\ref{Hubble})). It is reflected on one dimensional PDFs for $\\alpha$ which turned out to be flat meaning that the power of the present supernovae data to discriminate between various Generalized Chaplygin Gas models (differing by $\\alpha$) is weak. However, we argue that with future SNAP data it would be possible to differentiate between models with various value of $\\alpha$ parameter. Residual plots indicate the differences between $\\Lambda$CDM and Generalized Chaplygin Gas cosmologies at high redshifts. Therefore one can expect that future supernova experiments (e.g. SNAP) having access to higher redshifts will eventually resolve the issue whether the dark energy content of the Universe could be described as a Chaplygin gas. The discriminative power of forthcoming SNAP data has been illustrated on respective figures (Fig.26-29) obtained from the analysis on simulated SNAP data. \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f21.eps} \\caption{The density distribution (one dimensional PDF) for $\\Omega_{Ch}$ obtained from sample K3 by marginalization over remaining parameters of the model. We obtain the limit $\\Omega_{Ch} \\epsilon (0.61, 1.79)$ at the confidence level $95.4 \\%$. (Non-flat GCG model.)} \\label{fig:21} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f22.eps} \\caption{The density distribution (one dimensional PDF) for $A_0$ obtained from sample K3 by marginalization over remaining parameters of the model. We obtain the limit $A_0 \\epsilon (0.73,1)$ at the confidence level $95.4 \\%$.(Non-flat GCG model.)} \\label{fig:22} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f23.eps} \\caption{The density distribution (one dimensional PDF) for $\\alpha$ obtained from sample K3 by marginalization over remaining parameters of the model. We obtain the limit $\\alpha < 0.95$ at the confidence level $95.4 \\%$.(Non-flat GCG model.)} \\label{fig:23} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f24.eps} \\caption{The density distribution (one dimensional PDF) for $\\Omega_{k}$ obtained from sample K3 by marginalization over remaining parameters of the model. We obtain the limit $\\Omega_{k} \\epsilon (-1,0.23)$ at the confidence level $95.4 \\%$. (Non-flat GCG model.)} \\label{fig:24} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f25.eps} \\caption{The density distribution (one dimensional PDF) for $\\Omega_{m}$ obtained from sample K3 by marginalization over remaining parameters of the model. We obtain the limit $\\Omega_{m} \\epsilon( 0,0.53)$ at the confidence level $95.4 \\%$. (Non-flat GCG model.)} \\label{fig:25} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f26.eps} \\caption{Confidence levels on the plane $(A_0,\\alpha)$ for sample X1 ($\\Lambda$CDM model) of simulated SNAP data, marginalized over $\\Omega_m$. The figure shows the ellipses of preferred values of $A_0$ and $\\alpha$.} \\label{fig:26} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f27.eps} \\caption{Confidence levels on the plane $(A_0,\\alpha)$ for sample X2 (Cardassian model) of simulated SNAP data, marginalized over $\\Omega_m$. The figure shows the ellipses of preferred values of $A_0$ and $\\alpha$.} \\label{fig:27} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f28.eps} \\caption{Confidence levels on the plane $(A_0,\\alpha)$ for sample X3a (Generalized Chaplygin Gas model) of simulated SNAP data ($\\Omega_m=0$, $A_0=0.85$, $\\alpha=1.$), marginalized over $\\Omega_m$. The figure shows the ellipses of preferred values of $A_0$ and $\\alpha$.} \\label{fig:28} \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{f29.eps} \\caption{Confidence levels on the plane $(A_0,\\alpha)$ for sample X3b (Generalized Chaplygin Gas model) of simulated SNAP data ($\\Omega_m=0$, $A_0=0.76$, $\\alpha=0.49$), marginalized over $\\Omega_m$. The figure shows the ellipses of preferred values of $A_0$ and $\\alpha$.} \\label{fig:29} \\end{figure}" }, "0403/astro-ph0403149_arXiv.txt": { "abstract": "We studied the X-ray and optical absorption properties of 13 Gamma Ray Burst afterglows observed by BeppoSAX. We found that X-ray absorption in addition to the Galactic one along the line of sight is highly statistically significant in the two cases with the best statistics (probability $>99.9\\%$). In three other cases the presence of X-ray absorption is marginally significant (probability $\\sim 97\\%$). Measured rest frame absorbing column densities of hydrogen, $N_H$, range from 0.1 to 10.0 $\\times10^{22}$ cm$^{-2}$ (at 90\\% confidence level) assuming a solar metal abundance. X-ray absorption may be common, although the quality of present data does not allow us to reach a firm conclusion. We found that the rest frame column densities derived from XMM and Chrandra data as quoted in the literature are in good agreement with the BeppoSAX estimated rest frame $N_H$ range, supporting our result. For the same GRB afterglow sample we evaluated the rest frame visual extinction $A_{Vr}$. We fitted the optical-NIR afterglow photometry with a power law model corrected at short wavelengths by four different extinction curves. By comparing X-ray absorptions and optical extinction, we found that if a Galactic-like dust grain size distribution is assumed, a dust to gas ratio lower than the one observed in the Galaxy is required by the data. A dust to gas ratio $\\sim$ 1/10 than the Galactic one, as in the Small Magellanic Cloud (SMC) environment, has been tested using the SMC extinction curve, which produces good agreement between the best fit $N_H$ and $A_{Vr}$. We note, however, that the best fit $N_H$ values have been obtained by assuming solar metal abundances, while the metallicity of the SMC ISM is $\\sim$ 1/8 the solar one (Pei 1992). If such low metallicity were assumed, the best fit $N_H$ values would be higher by a factor of $\\sim7$, providing a significant increase of the $\\chi^2$. Alternative scenarios to explain simultaneously the optical and X-ray data involve dust with grain size distributions biased toward large grains. Possible mechanisms that can bring to such a grain size distribution are discussed. ", "introduction": "Soon after the $\\gamma$-ray flash, optical and X-ray afterglows of Gamma Ray Bursts (GRB) are among the brightest sources in the sky at cosmological redshifts. More than thirty GRB redshifts have been measured to date and their distribution ranges from 0.168 (Greiner et al. 2003) to 4.5 (Andersen et al. 2000) with a median of z $\\sim 1.1$ (excluding GRB 980425 if at z=0.0085). Follow-up observations of the GRB localized by BeppoSAX, the IPN and by the RXTE and HETE2 satellites, show that tens of minutes after the GRB the optical afterglow can be as bright as R=14-16 mag; a few hours later it can still be as bright as R=17-19 mag. An exceptionally bright example is the case of GRB 030329 for which the optical afterglow reached R=12.7 mag at 1.5 hours from the GRB and it decreased down to R=19 mag after $\\sim 10$ days (e.g. Price et al. 2003; Stanek et al. 2003). The X-rays afterglow can be as bright as the Crab Nebula a few minutes after the GRB, while 5-8 hours later it can be as bright as a bright AGN, i.e. $\\approx$mCrabs (see Frontera et al. 2000, also Fiore et al. 2000 for an estimate of the GRB afterglows logN-logF in the range 0.5-2.0 keV, where fluxes are integrated from minutes up to hours after the GRB event). This opens up the perspective for gathering spectra of high quality of sources at cosmological redshifts, provided that the afterglow can be observed in such short time scales. Spectral studies of GRB afterglows can provide crucial data on the environments in which GRB occurs. This can give us both important constraints on the nature of the GRB progenitor and on the interstellar matter (ISM) in the GRB host galaxies. Candidate GRB progenitor include: 'collapsars' (Woosley 1993, Paczynski 1998; MacFadyen \\& Woosley 1999), 'supranovae' (Vietri \\& Stella 1998), mergers of two neutron stars or a neutron star and a black hole (e.g. Eichler et al. 1989; Paczynski 1990; Narayan, Paczynski \\& Piran 1992; Meszaros \\& Rees 1992). Multiwavelength afterglow observations suggest the association of ``long GRB'' (for which the duration of the prompt event peaks at $\\sim20$s, e.g. Norris, Scargle \\& Bonnell 2000) progenitor with massive stars. Some indications are, among others: i) the GRB locations relative to their host galaxy (Bloom, Kulkarni \\& Djorgovski 2002a); ii) the detection of iron lines in the X-ray afterglow spectrum of five GRB (Piro et al. 1999, 2000; Yoshida et al. 1999; Antonelli et al. 2000; Amati et al. 2000); iii) the presence of late re-brightening in the light curves of some afterglows, interpreted as the evidence of an underlying supernova (e.g. Bloom et al. 2002b); iv) the presence of supernova features emerging in the optical afterglow spectra of GRB 030329 (Stanek et al. 2003, Hjorth et al. 2003; Kawabata et al. 2003) and of GRB 021211 (Della Valle et al. 2003). Because of the short lifetimes of massive stars, it is likely that their collapse happens close to their star forming region, that is, close to a dense and dusty environment (e.g. Fryer, Woosley \\& Hartman 1999; Perna \\& Belczynski 2002a). Therefore, one important element in favor of massive star progenitor would be the evidence of strong dust extinction. Indeed, the non detection of the $\\sim 60\\%$ of X-ray afterglows at optical wavelengths (dark GRB), favors this hypothesis. However, the optical spectra of GRB afterglows, do not generally show strong reddening (Simon et al. 2001; Galama \\& Wijers 2001). Moreover, Lazzati, Covino \\& Ghisellini (2002) show that even if dark GRB are located in the innermost regions of Galactic-like molecular clouds, in several cases the corresponding visual extinction is not enough to hide their optical afterglows. Therefore, a more accurate study of the GRB environment is needed. Prompt observations of GRB afterglows offer a new and distinctive path for the study of the matter in the immediate surroundings of the GRB, r$\\sim$1-10~pc (Perna \\& Loeb 1998; B\\\"ottcher et al. 1999; Perna et al. 2002b; Perna \\& Lazzati 2002c; Fox et al. 2003) and in the GRB host galaxy (Ciardi \\& Loeb 2000; Fiore 2001; Bloom et al. 2002a; Savaglio, Fall \\& Fiore 2003) X-ray and optical-UV spectroscopy can tell us about gas density and ionization status, metal abundances, dust content and kinematics of the matter in the GRB environment. This can be done by using both emission features and absorption features (see e.g. Kumar \\& Narayan 2003; B\\\"ottcher et al. 1999; Piro et al. 2000; Ghisellini et al. 2002). Although absorption features are more difficult to study than emission lines, they carry with them unbeatable information, because they probe matter along a single beam, i.e. along the line of sight to the background beacon. This greatly reduces the complications related to the matter geometry and dynamics, which strongly affects emission line studies. Even when such emission or absorption features are not detected or are not resolved, X-ray spectra can provide useful information from measures of low energy cut-offs due to photoelectric absorption Moreover, GRB afterglows can be used to study the ISM of their host galaxies. High redshifts galaxies have so far been studied mainly via `Lyman-break galaxies' (e.g. Steidel et al. 1999), and via galaxies which happen to lie along the line of sight to bright quasars, notably the `Damped Lyman-alpha' systems (DLA, e.g. Pettini et al. 1997, 1999). However, probably neither of these methods gives an undistorted view of the bulk of the typical high redshifts galaxy. Lyman-break galaxies are characterized by pronounced star-formation and their inferred chemical abundances may be related to these regions rather than being representative of typical high z galaxies (see e.g. Steidel et al. 1999). DLA studies are likely to be biased against dusty systems, since these would hide the background quasars in color-based surveys. Furthermore, DLA are often identified with dwarf or Low Surface Brightness galaxies and therefore may not be representative of the full high z galaxy population. On the other hand, GRB seem to occur well within the main body of their host galaxies, not in the outer halos (Bloom et al. 2002a), therefore, GRB afterglows can provide a powerful, independent tool to study the ISM of high redshifts galaxies (Castro et al. 2003, Mirabal et al. 2002, Fiore 2001; Savaglio et al. 2003). The latter authors compared metal column density in GRB host galaxies with those of DLAs. They find that while the column densities of iron and magnesium are similar or slightly higher than that in DLAs, the column of zinc are significantly higher. Since iron and magnesium, unlike zinc, tend to be depleted in dust, this finding indicates that a significant fraction of these elements is in dust since not observed through absorption lines, thus GRB are probably probing denser and dustier regions, than DLAs. We present in this paper the results of a systematic spectral analysis of 13 GRB afterglows observed by BeppoSAX in X-rays, and by several ground based telescopes in optical-UV, aimed at measuring and constraining the X-ray absorption along the line of sight as well as the extinction in the optical-UV bands. The comparison of the results obtained in the two spectral band will provide constraints on the dust to gas ratio of the absorbing matter and on the dust properties, as seen from several hours to a few days from the GRB event. The paper is organized as follows: \\S 2 shows the BeppoSAX X-ray afterglow selected sample; \\S 3 presents the X-ray data reduction and spectra extraction procedures, along with a discussion on the systematics affecting the column density measurements; \\S 4 presents the optical photometry of 9 GRB afterglows of our sample with a detected optical counterpart; results are presented and discussed in \\S 5 and \\S 6 respectively. ", "conclusions": "" }, "0403/astro-ph0403463_arXiv.txt": { "abstract": "We examine the X-ray spectra and variability of the sample of X-ray sources with $L_{\\rm X} \\approx 10^{31} - 10^{33}$~\\ergsec\\ identified within the inner 9\\arcmin\\ of the Galaxy by \\citet{mun03}. Very few of the sources exhibit intra-day or inter-month variations. We find that the spectra of the point sources near the Galactic center are very hard between 2--8~keV, even after accounting for absorption. When modeled as power laws the median photon index is $\\Gamma = 0.7$, while when modeled as thermal plasma we can only obtain lower limits to the temperature of $kT > 8$~keV. The combined spectra of the point sources is similarly hard, with a photon index of $\\Gamma = 0.8$. Strong line emission is observed from low-ionization, He-like, and H-like Fe, both in the average spectra and in the brightest individual sources. The line ratios of the highly-ionized Fe in the average spectra are consistent with emission from a plasma in thermal equilibrium. This line emission is observed whether average spectra are examined as a function of the count rate from the source, or as a function of the hardness ratios of individual sources. This suggests that the hardness of the spectra may in fact be to due local absorption that partially-covers the X-ray emitting regions in the Galactic center systems. We suggest that most of these sources are intermediate polars, which (1) often exhibit hard spectra with prominent Fe lines, (2) rarely exhibit either flares on short time scales or changes in their mean X-ray flux on long time scales, and (3) are the most numerous hard X-ray sources with comparable luminosities in the Galaxy. ", "introduction": "} Recent deep \\chandra\\ observations of the inner 9\\arcmin\\ around the super-massive black hole at the Galactic center have revealed over 2000 individual point-like X-ray sources \\citep{mun03}. The sources have luminosities between $10^{31}$ and $10^{33}$~\\ergsec \\citep[for a distance of 8 kpc to the Galactic center; see][]{mcn00}, and thus they probably represent some combination of young stellar objects, Wolf-Rayet and early O stars, interacting binaries (RS CVns), cataclysmic variables (CVs), young pulsars, and black holes and neutron stars accreting from binary companions (low- and high-mass X-ray binaries; LMXBs and HMXBs). However, the spectra of the Galactic center sources are very hard in the energy range of 2--8 keV. Spectra that are similarly hard have only been observed previously from magnetically accreting CVs (mCVs) and HMXB pulsars. Moreover, seven of the hard sources exhibit X-ray modulations with periods between 300~s and 4.5~h, which also suggests that they are magnetized white dwarfs or neutron stars \\citep{mun03c}. These basic observations are a good first step toward determining the natures of the point sources. However, if their natures can be determined conclusively, the large number of sources in the field would make it possible to study two important pieces of astrophysics: (1) the history of star formation at the Galactic center, and (2) the physics of X-ray production in accreting stellar remnants. How stars form at the Galactic center is still a mystery, because the strong tidal forces and milliGauss magnetic fields there should prevent all but the most massive molecular clouds from collapsing. Nonetheless, it appears that star formation has occurred recently, because three massive stellar clusters younger than $10^{7}$ years old lie within $\\approx 30$~pc of the Galactic center: IRS 16, the Arches, and the Quintuplet \\citep{kra95,pau01,fig99}. However, it is still a matter of debate as to whether the star formation is continuous or episodic, and whether it occurs only in localized regions or is relatively uniform throughout the Galactic center. \\citet{fig03} addressed this question by modeling the evolution of the population of luminous infrared stars, and concluded that the star formation is probably continuous. The X-ray sources at the Galactic center could provide an additional, independent constraint on the star formation history there, because they should be dominated by accreting stellar remnants. The size of the sample of X-ray sources --- an order of magnitude larger than the numbers of known LMXBs, HMXBs, and magnetic CVs --- also makes it a valuable database for studying the physics of X-ray production. Several outstanding questions could be addressed with the current data. If the sample contains large numbers of magnetic CVs, it could be used to determine the duty cycle of bright accretion states \\citep[e.g.,][]{gs88} and the fraction of such systems that exhibit hard spectral components \\citep[e.g.][]{ram03}. If there is a significant number of neutron star HMXBs, it may be possible to determine whether material accreted at rates far below the Eddington limit can penetrate the neutron star's magnetosphere and reach its surface \\citep{neg00,cam02,orl03}. Finally, the large sample of Galactic center X-ray sources would be useful for identifying systems with unusual properties. Previous hard X-ray surveys of the Galactic plane have identified several slowly-rotating accreting neutron stars and white dwarfs \\citep{kin98,tor99,oos99,sak00,sug00}, magnetic CVs with extremely strong emission lines from He-like Fe \\citep{mis96,ish98,ter99}, and accreting stellar remnants with high intrinsic absorption \\citep{pat03,mg03,wal03}. These systems could represent resting points for stellar remnants that have not been observed previously, and are therefore important for calculating the formation rate of such remnants in the Galaxy. In this paper, we take a further step toward the above goals by using the properties of the X-ray emission from the point sources near the Galactic center \\citep{mun03} to constrain better their natures. In Sections 2.1--2.3, we examine the spectra of the point sources both individually and averaged together, in order to determine the temperatures of the emitting regions. In Section 2.4, we search for short-term variability, which is often seen from coronal X-ray sources, and long-term variability, which is common in some accreting X-ray sources. In Section~3, we compare the properties of the observed sources with those of known classes of X-ray source. Finally, in Section~4, we briefly explore the future prospects for definitively identifying the natures of these sources. ", "conclusions": "We have established that, on average, the X-ray sources detected in 626~ks of \\chandra\\ ACIS-I observations of the field around \\sgrastar\\ have hard, $\\Gamma < 1$ spectra with prominent emission from He-like Fe at 6.7~keV (Figure~\\ref{fig:psmod} and Table~\\ref{tab:psint}). They also generally do not vary by more than factors of a few on time scales of hours or months. The best candidates for these hard X-ray sources are intermediate polars, which represent the most luminous and spectrally hardest 5\\% of all CVs. Therefore, the Galactic center X-ray sources are likely to be only a sub-sample of a population of $\\sim 10^{4}$ CVs located near the Galactic center. Although a single population of sources may dominate the image, there are certainly many classes of objects present in smaller numbers in the field. Determining the numbers of rare objects is particularly important. For instance, the numbers of massive Wolf-Rayet and O stars and faint neutron star high-mass X-ray binaries can constrain the recent rate of massive star formation near the Galactic center, while the numbers of LMXBS provide direct tests of the validity of unusual pathways for binary stellar evolution. For this reason, we are carrying out deep infrared observations of the Galactic center to identify counterparts to the X-ray sources. These observations will be useful for distinguishing CVs from, for example, HMXBs and WR/O stars. At at a distance of 8~kpc and with an extinction of $A_K \\approx 5$ \\citep{td03}, CVs should have $K$ magnitudes of 22--25, and therefore would be among the faintest detectable sources at the Galactic center \\citep{war95,hoa02}. In contrast, HMXBs and WR/O stars should have $K$ magnitudes brighter than 15 \\citep{zom90,weg94} and will be very easy to detect. Therefore, the prospects for identifying the natures of the Galactic center X-ray sources are promising." }, "0403/astro-ph0403655_arXiv.txt": { "abstract": "We present an analysis of rotation measure (RM) fluctuations from the Test Region of the Southern Galactic Plane Survey (SGPS), along with emission measure (EM) fluctuations in the same field taken from the Southern H-Alpha Sky Survey Atlas. The structure function of RM fluctuations shows a relatively steep slope at small scales (1~--~5 arcmin), a break in slope to a flatter structure function at intermediate scales (5~--~60 arcmin), and a systematic variation of the strength of fluctuations as a function of position angle on the sky at the largest scales (60~--~200 arcmin). The structure function of EM fluctuations shows similar behavior, although the lower resolution of the data prevents detection of a possible break in the spectrum. We interpret the anisotropy in RM/EM structure on large scales as resulting from a large-scale gradient in electron density (and possibly magnetic field) across the region. The break in the slope of the RM structure function at scales of $\\sim5$~arcmin can be explained by contributions from two spatially distinct magneto-ionized screens, most likely in the Local and Carina spiral arms. The observed structure function then implies that the outer scale of RM fluctuations in these screens is $\\sim2$~pc. Such behavior is in striking contrast to the expectation that interstellar turbulence forms an unbroken spectrum from kpc down to AU scales. We conclude that we have identified an additional source of enhanced turbulence, injected on scales of a few pc, possibly seen only in the Galactic plane. The most likely source of such turbulence is individual H\\,{\\sc ii} regions from relatively low-mass stars, whose characteristic scale size is similar to the outer scale of turbulence inferred here. These sources may be the dominant source of density and velocity fluctuations in warm ionized gas in the Galactic plane. ", "introduction": "The evidence for the presence of turbulence in the interstellar medium (ISM) is overwhelming. Although some studies of the characteristics of this turbulence (viz.\\ the inner and outer scales, shape of the spectrum and power law spectral index) indicate the presence of standard incompressible Kolmogorov (1941) turbulence, many observations indicate different kinds of turbulence, drivers and environments. Armstrong, Rickett \\& Spangler (1995) compiled observations of (among others) interstellar scattering of pulsars and extragalactic sources, dispersion measures of pulsars and rotation measures (RM) of extragalactic sources in one power spectrum. The result was the so-called ``big power law in the sky'', a Kolmogorov-like power spectrum of electron density fluctuations over 12 orders of magnitude, from a fraction of an AU to kiloparsecs. Although the big power law in the sky is very well-determined on small scales (up to $\\sim 0.001$~pc), the extension towards larger scales is based only on two types of measurements: the RM of extragalactic sources, which gives an upper limit to the amount of structure in the electron density because the contribution of the magnetic field is unknown; and from velocity measurements of H\\,{\\sc i}, making assumptions about the connection between neutral and ionized material. The dominant source for the large-scale energy input in the Galaxy is generally assumed to be supernova explosions (Spitzer 1978; Vollmer \\& Beckert 2002; Korpi et al.\\ 1999). Other possible sources are superbubbles, massive \\HII\\ regions and massive stellar winds \\citep{nf96}, Galactic fountains, chimneys, or gravitational scattering by transient spiral waves (see Sellwood \\& Balbus (1999) and references therein), gravitational instabilities in a shearing disk (Elmegreen, Elmegreen \\& Leitner 2003), or magneto-rotational instabilities \\citep{mk04}. The big power law in the sky certainly suggests the classical scenario of turbulent energy input on large scales, which cascades down to smaller scales until the energy is dissipated on the smallest scale \\citep{k41}. However, there are indications of other types of turbulence in the ISM as well. \\cite{ms96} found a break in the structure function of RMs of extragalactic sources and emission measures (EM) of the warm ionized gas on scales of a few parsec, which they interpreted as a transition from three-dimensional to two-dimensional turbulence as one moves from small to large scales. Analytic theory (Goldreich \\& Sridhar 1995, 1997) and simulations (e.g.\\ Cho, Lazarian \\& Vishniac 2002; Maron \\& Goldreich 2001) show that magnetic fields can lead to anisotropic turbulence, which is predicted to exhibit the same power law spectral index as Kolmogorov turbulence. Furthermore, observations of RMs of extragalactic sources (Simonetti \\& Cordes 1986; Spangler \\& Reynolds 1990; Clegg \\etal\\ 1992) show higher amplitudes of structure in RM in the Galactic plane than out of the plane, suggesting the existence of an additional source of structure on parsec scales. Results from interstellar scattering show a similar enhancement of structure in the inner Galaxy, but on much smaller scales (Rao \\& Ananthakrisnan 1984; Dennison \\etal\\ 1984; Anantharamaiah \\& Narayan 1988). Structure in the ionized ISM can be studied by way of structure functions (SFs). Earlier determinations of SFs of RM using extragalactic sources or the diffuse synchrotron background yielded results that varied widely with scale and area in the sky. Flat structure functions (indicating no structure on the probed scales) found by Simonetti, Cordes \\& Spangler (1984) on scales larger than $4\\degr$ towards the Galactic pole were interpreted as showing structure intrinsic to the extragalactic sources with a negligible Galactic contribution. However, this cannot explain the flat SFs from low-latitude extragalactic sources found by \\citet{scs84} on scales $\\ga 4\\degr$ and by \\citet{ccs92} on scales $\\ga 1\\degr$. Furthermore, Sun \\& Han (2004) have recently found shallow SFs of RM in the Galactic plane, and a flat SF at the North Galactic Pole, also from extragalactic sources. Haverkorn, Katgert \\& de Bruyn (2003a) studied the structure function of RM from diffuse radio emission, and found very shallow slopes for two fields at intermediate latitudes. In this paper we study the turbulent structure in ionized gas in the inner Galactic plane, by means of SFs of RM from the diffuse synchrotron background, and of SFs of EM from H$\\alpha$ emission, both in a region in the inner Galactic plane. In Section~\\ref{s:obs} we present our radio polarimetric observations of the Galactic synchrotron background in the Galactic plane. Section~\\ref{s:sfdet} discusses the computation of the structure function, while in Section~\\ref{s:sfint} we interpret the structure function as arising from Faraday screens in two spiral arms along the line of sight, both exhibiting turbulent structure. In Section~\\ref{s:hii}, we discuss evidence for enhanced density fluctuations in the Galactic plane, and propose that the structure is dominated by discrete \\HII\\ regions. ", "conclusions": "\\label{s:con} We have analyzed structure functions (SF) of RM and EM data in the SGPS Test Region. These show very consistent results: \\begin{itemize} \\item the SFs of both RM and EM, \\sfrm\\ and \\sfem, exhibit a linear slope in log-log space, with \\sfrm\\ showing a break at $\\sim4\\arcmin$ to a flatter slope at larger angular scales; \\item the slope at angular scales $\\ga4\\arcmin$ is shallow though non-zero in both \\sfrm\\ and \\sfem, with a SF spectral index $b \\approx 0.2$; \\item at scales larger than about a degree, the spectral index of \\sfrm\\ and \\sfem\\ is anisotropic, forming a quasi-sinusoidal dependence on position angle. \\end{itemize} The anisotropy in one-dimensional SFs at large scales is explained by a large-scale gradient in electron density, possibly accompanied by magnetic field, across the field of view due to a foreground structure. The break in \\sfrm\\ and shallowness of the slope in both \\sfrm\\ and \\sfem\\ at smaller scales can be explained by the superposition of two contributions to EM and RM of the Local and Carina spiral arms. Within such a model, we infer the outer scale of structure in the spiral arms to be about 2~pc. As Kolmogorov-like turbulence is observed in the ISM on scales much larger than a few pc, these results constitute evidence for an additional contribution to turbulent fluctuations in the Galactic plane. The inferred outer scale of fluctuations agrees with the size of Str\\\"omgren spheres around the most abundant late-type B~stars. Therefore, \\HII\\ regions could provide the dominant source of structure on pc scales in the Galactic spiral arms. The suggestion that the break in \\sfrm\\ and the shallowness of both \\sfrm\\ and \\sfem\\ is due to two separate Faraday screens will be tested in a forthcoming paper by studying SFs in different regions of the sky. This can also shed light on the occurrence of any spatial variations of the turbulent spectrum. Furthermore, the combined RM and EM data should enable at least partial decoupling of magnetic field and electron density, and the amplitudes of the SF will yield information on this additional component of structure in the ionized ISM in the Galactic plane." }, "0403/astro-ph0403525_arXiv.txt": { "abstract": "We present deep CCD photometry in the $VI$ passbands using the WIYN 3.5m telescope of a field located approximately 20' southeast of the center of M33; this field includes the region studied by Mould \\& Kristian in their 1986 paper. The color-magnitude diagram (CMD) extends to I$\\sim$25 and shows a prominent red giant branch (RGB), along with significant numbers of asymptotic giant branch and young main sequence stars. The red clump of core helium burning stars is also discernable near the limit of our CMD. The I-band apparent magnitude of the red giant branch tip implies a distance modulus of $(m-M)_I = 24.77 \\pm 0.06$, which combined with an adopted reddening of $E(V-I)=0.06\\pm0.02$ yields an absolute modulus of $(m-M)_0 = 24.69 \\pm 0.07$ (867$\\pm$28 kpc) for M33. Over the range of deprojected radii covered by our field ($\\sim$8.5 to $\\sim$12.5 kpc), we find a significant age gradient with an upper limit of $\\sim$1 Gyr ($\\sim$0.25 Gyr/kpc). Comparison of the RGB photometry to empirical giant branch sequences for Galactic globulars allows us to use the dereddened color of these stars to construct a metallicity distribution function (MDF). The primary peak in the MDF is at a metallicity of $[Fe/H]$$\\sim$--1.0 with a tail to lower abundances. The peak does show radial variation with a slope of $\\Delta$[Fe/H]/$\\Delta$$R_{deproj}$ = --0.06 $\\pm$ 0.01 dex/kpc. This gradient is consistent with the variation seen in the inner {\\it disk} regions of M33. As such, we conclude that the vast majority of stars in this field belong to the disk of M33, not the halo as previously thought. ", "introduction": "In recent years M33 has been the target of many systematic studies. From searches for variable stars to pin down one of the fundamental distance estimates in the Cepheid distance latter (e.g. Macri et al. 2001), to kinematical studies of the stellar populations (e.g. Chandar et al. 2002), to X-ray surveys (e.g. Haberl \\& Pietsch 2001). However, in spite of M33 being the second closest spiral galaxy after M31, close enough for the brighter members of its stellar population to be resolved, the properties of its field halo stars have received comparatively little attention. The first attempt to study the field halo stars in M33 using CCDs was that of Mould \\& Kristian (1986, hereafter MK86). In this classic paper, the authors observed an approximately 5x5 arcmin field (see Fig. 1) with the Palomar 5m telescope along with one of the first generation science-grade CCD detectors, a Texas Instruments 800 x 800 pixel array. Their color-magnitude diagram (CMD), based on aperture photometry of 215 stars transformed onto the Cousins system (Cousins 1976a,b), extends from above the first ascent red giant branch (RGB) tip down to $\\sim$1.5 mags below the tip (I$\\sim$22.5). From this diagram, they drew two primary conclusions. First, by comparing the M33 RGB with those of the Galactic globular clusters M92 and 47 Tuc, they found the mean metal abundance of the field stars to be $\\langle$[M/H]$\\rangle$ = --2.2 $\\pm$ 0.8. This result was rather surprising given the much higher metallicity ($\\langle$[M/H]$\\rangle > -0.8$) they found for the halo of M31. Second, from the bolometric magnitude of the RGB tip and an adopted reddening of $E(V-I) = 0.06$, they calculated an absolute distance modulus of 24.8 $\\pm$ 0.2 for M33. Cuillandre, Lequeux, \\& Lionard (1999) used the UH8k CCD camera at the prime focus of the CFHT 3.6m to image a 28 x 28 arcmin field in a region of M33 that included the MK86 field. Their $(V,V-I)$ CMD reaches as faint as V $\\sim$ 25.5. Cuillandre et al. (1999) adopt an M33 distance modulus of $(m-M)_0 = 24.82$ and estimate the line-of-sight reddening ($E(V-I) = 0.08$) by averaging values based on the 21-cm line (Hartmann 1994) and from foreground stars (Johnson \\& Joner 1987). Based on these values, a comparison with the RGB sequences of Da Costa \\& Armandroff (1990) yields a mean metal abundance of [Fe/H] = --1.0 with a metallicity spread from $\\sim$--1.5 to $\\sim$--0.6 dex for the field halo stars in M33. This mean abundance is clearly much higher than the value advocated by MK86 and a few tenths of a dex higher than the peak abundance of Milky Way field halo stars (e.g. Norris 1994; Carney et al. 1996;). Most recently, Davidge (2003) has used the Gemini North 8m telescope equipped with GMOS to image a field located $\\sim$9 kpc in projected distance from M33 ($\\sim$15.5 kpc deprojected distance), approximately twice as far out as the MK86 field. The g', r', i', z' CMDs of Davidge (2003) are based on a field of view of only 5x5 arcmin and are not extraordinarily deep, reaching some 2 magnitudes below the tip of the RGB, but they are notable as being the first such M33 data at these extreme galactocentric distances. Davidge's conclusions are twofold. First, based on the color of the upper RGB, Davidge (2003) determines a mean abundance of [Fe/H] = $-1.0 \\pm 0.3$ (random) $\\pm 0.3$ (systematic) for this region of M33. Second, the presence of a significant number of bright asymptotic giant branch stars suggests that an intermediate-age population exists outside of the `young star-forming' disk of M33. This latter result is consistent with the work of several investigators (e.g. Sarajedini et al. 1998;2000, Chandar et al. 2002) who have suggested that the halo star clusters in M33 are several Gyr younger than similar clusters in the Milky Way's halo. Within this context, the present series of papers attempts to shed further light on the field halo stellar population(s) of M33. This first paper presents deep ground-based photometry of the MK86 field along with an analysis designed to probe the metallicity distribution function of this region of M33. We begin with a discussion of the observational material in the next section along with details of the photometric reduction procedure. Section 3 describes our artificial star experiments and how they are used to gauge the photometric completeness and errors. We compare our photometry to that of MK86 in Sec. 4. The analysis portion of the paper begins in Sec. 5 in which we use the luminosity function of the RGB to estimate the distance to M33. In Sec. 6, the radial variations of the stellar populations present in our field are analyzed. Section 7 presents the metallicity distribution function and what it implies for this region of M33. Finally, our results are summarized in Sec. 8. ", "conclusions": "In this work, we present deep $VI$ photometry of a field in M33, that includes the region studied by Mould \\& Kristian(1986), observed with the WIYN 3.5m telescope. Based on the CMD, which reaches past the helium burning red clump, we draw the following conclusions. \\noindent 1) The I-band apparent magnitude of the TRGB is measured to be $\\rm I_{\\rm TRGB} = 20.75 \\pm 0.04$, where the error includes the measurement error (0.02 mag) along with random (0.02 mag) and systematic (0.03 mag) errors in the photometry. \\noindent 2) The I-band apparent magnitude of the TRGB implies a distance modulus of $(m-M)_I = 24.77 \\pm 0.06$, which combined with an adopted reddening of $E(V-I)=0.06\\pm0.02$ yields an absolute modulus of $(m-M)_0 = 24.69 \\pm 0.07$ (867$\\pm$28 kpc) for M33. \\noindent 3) By examining the spatial properties of stars in various parts of the CMD, we find that, over the range of deprojected radii covered by our field ($\\sim$8.5 to $\\sim$12.5 kpc), there is a significant age gradient with an upper limit of $\\sim$1 Gyr (i.e., $\\sim$0.25 Gyr/kpc). \\noindent 4) Using the empirical RGB grid constructed by Saviane et al. (2000), we have converted the dereddened color of each RGB star to a metallicity. The resultant metallicity distribution function (MDF) displays a primary peak at a metallicity of $[Fe/H]$$\\sim$--1.0 with a tail to lower abundances. The peak shows a radial variation with a slope of $\\Delta$[Fe/H]/$\\Delta$$R_{deproj}$ = --0.06 $\\pm$ 0.01 dex/kpc. This gradient is consistent with the variation seen in the inner {\\it disk} regions of M33. Therefore, we conclude that the vast majority of stars in this field belong to the disk of M33, not the halo as previously thought." }, "0403/astro-ph0403239_arXiv.txt": { "abstract": "State-of-the-art NLTE model-atmosphere codes have arrived at a high level of numerical sophistication and are now useful tools to analyze high-quality spectra from the infrared to the X-ray wavelength range. The capacity of current computers permit calculations which include line spectra from all elements from hydrogen up to the iron group. The lack of reliable atomic data has become a critical problem for further progress. We summarize available sources of atomic data, and discuss how these are implemented in the T\\\"ubingen Model-Atmosphere Package \\linebreak \\textsc{tmap}. We describe our Iron Opacity Interface {\\sc IrOnIc} which is used to calculate opacities of iron-group elements from Kurucz's and the Opacity Project's data. We propose general use of the T\\\"ubingen Model-Atom Database \\linebreak {\\sc tmad}, which would allow an easy exchange of ready-to-use model atoms between all model-atmosphere groups. The comparison of model-atmo\\-sphere calculations would then be much easier, and would save a great deal of manpower that is presently consumed preparing suitable model atoms for spectral analyses. ", "introduction": "In the early 80's of the last century, the implementation of approximate lambda-operators (ALO, leading to ``accelerated lambda iteration'', ALI) in the NLTE model-atmosphere codes at Kiel by Werner \\& Husfeld (1985) and Werner (1986) provided an efficient method to calculate synthetic stellar spectra of hot stars. Together with the access to the fourth {\\sc cray} computer (a {\\sc cray-1\\,m}) in Germany (at the Konrad-Zuse-Zentrum f\\\"ur Informationstechnik Berlin), which was installed in February 1984, powerful tools for spectral analysis of hot star spectra were developed. At the end of 1987, the Rechenzentrum der Universit\\\"at Kiel installed a {\\sc cray\\,x-mp}. Access to this machine (in the framework of the Norddeutscher Vektorrechner-Verbund) and the following, even more powerful, {\\sc cray} computers, made all our efforts possible. The first of our NLTE codes, \\textsc{hymoc} ({\\sc HYd}rogen {\\sc MO}del-atmosphere {\\sc C}ode) (Werner 1986), was able to calculate only pure-hydrogen model atmospheres. However, inasmuch as the atomic data for hydrogen are well known, it was an ideal tool to investigate the numerical approximations, and limitations to the size of model atoms, used in earlier calculations, which were necessary before the ALI method was developed (Rauch \\& Werner 1988). The next code \\textsc{pro2} ({\\sc pro}gram no.\\,2) (Werner 1988; Werner \\& Dreizler 1999) is much more flexible, and is able to take into account all elements up to the iron group (Dreizler \\& Werner 1993; Rauch 1997). It has been used successfully for the analysis of hot stars (e.g.~Rauch \\& Werner 1991; Rauch 1993; Rauch 2000). Our NLTE group moved from Bamberg (1993) and Potsdam (1995) to T\\\"ubingen (since 1996) where \\textsc{tmap}, the state-of-the-art {\\sc T}\\\"ubingen {\\sc M}odel {\\sc A}tmosphere {\\sc P}ackage was created. With the newly developed \\textsc{IrOnIc} code (\\S 4), it was possible to calculate an extended grid of realistic stellar fluxes from models which take into account the opacities from H -- Ni (Rauch \\& Deetjen 2001). These models are plane-parallel, in hydrostatic and radiative equilibrium, have 350 atomic levels which are treated in NLTE, about 1,000 individual lines from H - Ca, and millions of lines from the iron-group elements. The state of the field of spectral-analysis of hot stars has completely changed within the last two decades. At the beginning of the 80's, the main obstacles were insufficient numerical methods and computational capacities. Rauch \\& Werner (1991) have shown the enormous progress which came with the ALI method, with examples of very detailed H + He + C + N models in contrast to ``classical'' H + He models. At present, the lack of reliable atomic data for metals, line-broadening tables, etc.~set undesirable limits to highly-developed NLTE codes. This lack often hampers an adequate analysis e.g.~of high-resolution UV spectra provided by the STIS (Space Telescope Imaging Spectrograph) aboard the Hubble Space Telescope (HST) or by the Far Ultraviolet Spectroscopic Explorer (FUSE). ", "conclusions": "Present NLTE model-atmosphere codes have reached a very high level of sophistication. Now, strong efforts to achieve adequate atomic data must be continued in order to be able to analyze reliably the high-quality spectra which are already available from the infrared to the X-ray range. {\\it Everyone who is calculating stellar atmospheres should be aware of the important role that atomic data plays, because one thing is for sure -- even if you use a perfect code, if you put rubbish in, you will get rubbish out.}" }, "0403/astro-ph0403713_arXiv.txt": { "abstract": "High ambient interstellar pressure is suggested as a possible factor to explain the ubiquitous observed growth-rate discrepancy for supernova-driven superbubbles and stellar wind bubbles. Pressures of $P/k\\sim 10^5\\ \\ccK$ are plausible for regions with high star formation rates, and these values are intermediate between the estimated Galactic mid-plane pressure and those observed in starburst galaxies. High-pressure components also are commonly seen in Galactic ISM localizations. We demonstrate the sensitivity of shell growth to the ambient pressure, and suggest that superbubbles ultimately might serve as ISM barometers. ", "introduction": "Mechanical feedback from supernovae (SNe) and stellar winds is an important driver of galactic evolutionary processes. It affects the phase balance and physical conditions of the interstellar medium (ISM), which in turn determine star-forming conditions, galactic chemical evolution, and properties of the intergalactic medium. The standard paradigm for mechanical feedback is based on the model for adiabatic evolution of the shells and superbubbles (e.g., Pikel'ner 1968; Weaver {\\etal}1977; Mac Low \\& McCray 1988) that are the direct consequence of SN and stellar wind action. While this model for pressure-driven superbubbles is broadly consistent with observations spanning scales from individual stellar wind bubbles to galactic superwinds, nagging problems persist in comparisons with observations (e.g., see Oey 2004 for a review). Specifics regarding the energy budgets, fate of the shock-heated $10^6$ K gas, and later-stage evolution are lacking and have profound consequences for galactic evolution. One problem that is empirically well-established is the result that most superbubbles apparently grow more slowly than expected. This has been observed in individual stellar wind bubbles such as Wolf-Rayet nebulae (Treffers \\& Chu 1982; Garc\\'\\i a-Segura \\& Mac Low 1995; Drissen {\\etal}1995), as well as in superbubbles powered by OB associations (e.g., Oey 1996$a$; Oey \\& Smedley 1998; Brown {\\etal}1995; Saken {\\etal}1992). This growth-rate discrepancy has been identified in young, nebular shell systems, in which the parent OB association is still present; thus the input mechanical power is well-constrained. The discrepancy is seen both in objects that show no evidence of previous supernova activity, and in ones where one or two supernovae have already exploded (Oey 1996$a$; hereafter O96). For constant input mechanical power $L$ and ambient number density $n$, the evolution of the shell radius is given by (e.g., Castor, McCray, \\& Weaver 1975), \\begin{equation}\\label{eqR} R = 68.9\\ (L_{38}/n)^{1/5}\\ t_6^{3/5}\\ \\ \\rm pc \\quad , \\end{equation} where $L_{38}$ is $L$ in units of $10^{38}\\ \\ergs$, and $t_6$ is age of the bubble in Myr. The shell expansion velocity $v$ is the time derivative of equation~\\ref{eqR}. One possible solution to the growth-rate discrepancy suggests that the input parameter $L/n$ is systematically overestimated. For eight nebular superbubbles with well-constrained $R$, $v$, $L$, and $t$, Oey and collaborators (O96; Oey \\& Massey 1995; Oey \\& Smedley 1998) showed that $L/n$ would need to be reduced by a factor of several, perhaps up to an order of magnitude, to reconcile the observations with prediction. Since stellar wind power $L$ is sensitive to the stellar mass, a substantial uncertainty in $L$ is not unreasonable. As shown by multi-wavelength observations of three of the superbubbles (Oey {\\etal}2002), the multi-phase ambient ISM also renders $n$ similarly uncertain. However, the implication of a {\\it systematic} growth-rate discrepancy remains difficult to explain. Another favorite candidate to solve the problem is cooling of the hot interior whose pressure drives the shell growth. If this scenario is correct, it implies a departure from the adiabatic evolution. The mass within the hot bubble interior is dominated by material evaporated from the cool shell walls, and could be supplemented by additional material evaporated and ablated from small clouds that are overrun by the expanding shocks (Cowie \\& McKee 1977; McKee {\\etal}1984; Arthur \\& Henney 1996). This enhanced interior density would facilitate radiative cooling. Silich {\\etal}(2001) and Silich \\& Oey (2002) also show that the enhanced metallicity caused by SN explosions and stellar products can further facilitate the cooling, especially for low-metallicity objects. However, increased X-ray luminosities that are expected from enhanced cooling thus far have not been observed (Chu {\\etal}2003; Chu {\\etal}1995). The superbubbles studied by Oey and collaborators are all located in the Large Magellanic Cloud (LMC). For these objects, Silich \\& Franco (1999) suggested that the ambient environment and viewing geometry conspire to yield misleading observed shell dynamics. They suggest that the superbubbles are more extended perpendicular to the galaxy's plane, as would be expected in the plane-stratified density distribution of disk galaxies (but see also Maciejewski \\& Cox 1999). The elongation of the shells would not be apparent because of the LMC's almost face-on orientation. While this is an attractive suggestion for the LMC objects, it does not explain the growth-rate discrepancy seen in Galactic (e.g., Brown {\\etal}1995; Saken {\\etal}1992) and M33 (Hunter {\\etal}1995) objects. Nevertheless, it is apparent that the ambient environment plays a crucial role in the superbubble growth and evolution. In addition to the work of Silich \\& Franco (1999), other studies have shown that the shell dynamics are sensitive to the ambient density structure (e.g., Oey \\& Smedley 1998; Mac Low {\\etal}1998). Multi-wavelength observations also show that the ambient multiphase gas distribution is difficult to constrain without direct such observations (Oey {\\etal}2002). ", "conclusions": "Our models clearly show that increasing the ambient interstellar pressure by an order of magnitude, from $P_e/k = 1\\times 10^4$ to $1\\times 10^5\\ \\ccK$, can impede the shell growth to a degree that could fully account for the observed growth-rate discrepancy. In \\S 2, we presented arguments that such high interstellar pressures could exist, especially based on the dependence of $P_e$ on star-formation rate. While other factors mentioned in \\S 1, namely, overestimated $L/n$, elevated radiative cooling, and viewing geometry, could all be additional factors that contribute to the growth-rate discrepancy, we note that the multi-phase gas morphology is more consistent with high interstellar pressure dominating this effect. Finally, as noted by Oey \\& Clarke (1997), the assumed global value of $P_e$ plays a critical role in determining the characteristic final sizes of old, SN-dominated superbubbles, and hence, the superbubble size distribution, which is dominated by pressure-confined shells. This, in turn, determines the interstellar porosity and filling factor of the hot, ionized medium in star-forming galaxies. With adequate clarification in the superbubble evolution process and input parameters, the superbubble sizes, kinematics, and morphologies could potentially provide barometers for the interstellar pressure. These diagnostics could be especially useful in other galaxies, which have fewer available empirical pressure indicators than the Milky Way." }, "0403/hep-ph0403004_arXiv.txt": { "abstract": "\\noindent Assuming that cosmological dark matter consists of weakly interacting massive particles, we use the recent precise measurement of cosmological parameters to predict the guaranteed rates of production of such particles in association with photons at electron-positron colliders. Our approach is based on general physical principles such as detailed balancing and soft/collinear factorization. It leads to predictions that are valid across a broad range of models containing WIMPs, including supersymmetry, universal extra dimensions, and many others. We also discuss the discovery prospects for the predicted experimental signatures. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403707_arXiv.txt": { "abstract": "{It is commonly adopted that X-rays from O stars are produced deep inside the stellar wind, and transported outwards through the bulk of the expanding matter which attenuates the radiation and affects the shape of emission line profiles. The ability of the X-ray observatories Chandra and XMM-Newton to resolve these lines spectroscopically provided a stringent test for the theory of the X-ray production. It turned out that none of the existing models was able to fit the observations consistently. The possible caveat of these models was the underlying assumption of a smooth stellar wind. Motivated by the various evidences that the stellar winds are in fact structured, we present a 2-D numerical model of a stochastic, inhomogeneous wind. Small parcels of hot, X-ray emitting gas are permeated with cool, absorbing wind material which is compressed into thin shell fragments. Wind fragmentation alters the radiative transfer drastically, compared to homogeneous models of the same mass-loss rate. X-rays produced deep inside the wind, which would be totally absorbed in a homogeneous flow, can effectively escape from a fragmented wind. The wind absorption becomes wavelength independent if the individual fragments are optically thick. The X-ray line profiles are flat-topped in the blue part and decline steeply in the red part for the winds with short acceleration zone. For the winds where the acceleration extends over significant distances, the lines can appear nearly symmetric and only slightly blueshifted, in contrast to the skewed, triangular line profiles typically obtained from homogeneous wind models of high optical depth. We show that profiles from a fragmented wind model can reproduce the observed line profiles from $\\zeta$\\,Orionis. The present numerical modeling confirms the results from a previous study, where we derived analytical formulae from a statistical treatment. ", "introduction": "Hot massive stars possess strong stellar winds, as discovered by the advent of UV spectroscopy (Morton \\cite{Morton67}). One decade later it was detected that these stars emit X-rays (Harnden et al.\\ \\cite{Harnden79}, Seward et al.\\ \\cite{Seward79}). The origin of these X-rays is debated controversially. One of the first suggestions was the existence of a hot corona close to the stellar photosphere. However, the surrounding stellar wind should imprint strong K-shell absorption edges on the X-ray spectrum, which were not observed (Cassinelli \\& Swank \\cite{CS83}). The base-coronal model was finally ruled out because of the lack of characteristic coronal emission lines (Baade \\& Lucy \\cite{Lucy82}). Therefore it became commonly adopted that X-rays from hot stars are produced within their stellar winds, although most of the wind gas is obviously ``cool''. The driving mechanism for the mass-loss from OB stars has been identified with radiation pressure on spectral lines. A corresponding theory was developed by Castor et al.\\ (\\cite{CAK75}, ``CAK''). After a number of improvements, this theory seems to predict the observed mass-loss rates and wind velocities correctly (Pauldrach et al.\\ \\cite{Pauldrach86}). However, it was pointed out early (Lucy \\& Solomon \\cite{LucySolomon70}), and later further investigated (Carlberg \\cite{Carlberg80}; Owocki \\& Rybicki \\cite{OR84}), that the stationary solution for a line-driven wind is unstable; small perturbation should grow quickly and form shocks. This ``de-shadowing instability'' is implicitly suppressed in the CAK theory. The most detailed hydrodynamic modeling of this line driven instability was presented by Feldmeier et al.\\ (\\cite{AF97}). These calculations show how initially small perturbations grow and form strong shocks that emit X-rays. This ``wind shock model'' was principally able to reproduce the X-ray flux from the O supergiant $\\zeta$~Ori\\,A, which had been observed at that time only at low spectral resolution. The ultimate test for models of X-ray production in stellar winds became possible with the launch of the X-ray observatories Chandra and XMM-Newton. Their spectrographs allow for the first time to resolve X-ray emission line profiles. Surprisingly, the shape of the X-ray line profiles turned out to be strikingly different among the few individual O stars observed so far: \\begin{description} \\item {$\\theta^1$~Ori\\,C (Schulz et al.\\ \\cite{Schulz00}) and $\\zeta$~Ori\\,A (Waldron \\& Cassinelli \\cite{WaldCass01})} show profiles which appear wind-broadened, symmetric and unshifted with respect to the center wavelength; \\item {$\\delta$~Ori\\,A (Miller et al.\\ \\cite{Milleretal02})} shows profiles which are narrower, symmetric, and also centered at the laboratory wavelength; \\item {$\\zeta$~Pup (Cassinelli et al.\\ \\cite{Cass01}; Kahn et al.\\ \\cite{Kahnetal01})} shows strongly broadened profiles which are blueshifted against the rest wavelength but otherwise appear to be symmetric. \\end{description} If X-rays are produced deep inside the stellar wind, they have to propagate through the absorbing cool wind before they can emerge towards the observer. The opacity at X-rays energies is much larger than in the visual and UV, due to the K-shell absorption of abundant metal ions and the He\\,{\\sc ii} edge. Red-shifted photons of an emission line originate in the back hemisphere of the wind, and suffer much stronger absorption than photons in the blueshifted part of the line profile. Hence, as soon as the absorbing wind has a significant optical depth, the line profiles should be blueshifted and skewed, which means their shape resembles that of a triangle, with the flux maximum at maximum blueshift (MacFarlane et al.\\ \\cite{MacFarlane91}). None of the observed stars meets this prediction. Kramer et al.\\ (\\cite{KCO03}) were able to fit eight X-ray lines observed in $\\zeta$\\,Puppis with a homogeneous wind model. However, these fits could only be achieved by assuming much less absorption in the wind than expected from the generally adopted mass-loss rate of this star. Moreover, Kramer et al.\\ (\\cite{KCO03}) state that there is not a strong trend of absorption with wavelength, in contrast to what is expected from the energy dependence of the mass absorption coefficient. Thus even for the prototype single O star $\\zeta$~Puppis the wind shock model fails to reproduce observed X-ray emission lines consistently. At this point the question arose whether this model is adequate. We emphasize that the whole interpretation of X-ray spectra had been based so far on over-simplified models, assuming a {\\em homogeneous} distribution of the absorbing material. By contrast, there is strong empirical evidence that stellar winds are in fact strongly clumped (e.g.\\ Hamann \\& Koesterke \\cite{HK98}, Eversberg et al.\\ \\cite{Ev98}, Puls et al.\\ \\cite{Puls03}). This supports hydrodynamic models (Owocki et al.\\ \\cite{OCR88}, Feldmeier \\cite{AF95}) which show that most of the material in a radiatively driven wind is compressed into a series of dense shells by the instability. As the hydrodynamic models are one-dimensional, they cannot tell anything about the lateral structure of these wind inhomogeneities. The de-shadowing instability acts only in the radial direction; on the other hand, there is no obvious mechanism which could synchronize the wind in different directions. The observed variability in wind lines is not very pronounced, suggesting that the dense shells break up into a large number of fragments. Wind inhomogeneity alters the radiative transfer significantly. We have studied the effects of wind fragmentation on the X-ray line formation in Feldmeier et al.\\ (\\cite{PaperI}, Paper\\,I), using a statistical approach that holds in the limit of infinitely many fragments. We showed analytically that the continuum wind opacity is greatly reduced in a structured wind, compared to a smooth wind with the same mass-loss rate. The line profiles we obtained are broad, blueshifted and flat-topped, i.e. symmetric, and thus of promising similarity to those observed in $\\zeta$~Pup. In the present paper we construct a stochastic wind model using a numerical approach. This allows to drop a number of idealizations in favor of a more realistic description. \\begin{description} \\item {\\it Statistical approach.} While Paper\\,I relied on the statistical limit of very many fragments, we allow now for an arbitrary scale of fragmentation. \\item {\\it Wind geometry.} We study models with different spatial distributions of fragments, both in the radial and lateral direction. \\item {\\it Wind velocity law.} We abandon the restriction to a constant wind velocity and allow for so-called $\\beta$-velocity laws. \\end{description} Last not least, another purpose of the present paper is to better understand some surprising results obtained in Paper\\,I. Despite of our focus on X-ray emission line formation, we want to emphasize their general importance for radiative transfer in inhomogeneous media. In the next section we will introduce the stochastic wind model. Section\\,3 provides the formalism for solving the radiative transfer in a fragmented wind, as implemented in our numerical modeling. The basic effects of structured absorption on line profiles are illustrated in Sect.\\,4 using simplified examples. In Sect.\\,5 we introduce the concept of an effective opacity, and compare the line profiles obtained numerically from the stochastic wind model with the results from an analytical, statistical treatment. In Sect.\\,6 we evaluate line profiles using realistic assumptions and explore the parameter space. Conclusions are drawn in the final Sect.\\,7. A forthcoming paper will be devoted to detailed fits of observations. ", "conclusions": "A main motivation for our fragmented wind model is to explain the observed X-ray line profiles from O stars, which cannot be reproduced so far. We may check now whether we came closer to this goal. Figure\\,\\ref{fig:zeta_Ori} shows the profile of the Ne\\,{\\sc x} line from $\\zeta$\\,Orionis, as observed with the Chandra satellite (histogram). The smooth line is a tentative comparison with one of our calculations. From the known mass-loss rate of the star, the continuum at the frequency of this line may be optically thick (Waldron et al.\\ \\cite{WaldCass01}). Therefore we select a suitable profile (with $\\beta = 2$) from the set in Fig.\\,\\ref{fig:bvthick} with $\\tau_\\ast = 10$. The dimensionless frequency is converted into wavelength according to the terminal wind velocity of $\\zeta$\\,Orionis (2100\\,km\\,s$^{-1}$), and the profile is convolved with a Gaussian of 0.023\\,\\AA\\ FWHM to account for instrumental broadening. The good agreement just demonstrates that our fragmented wind model can reproduce emergent line profiles of the observed shape, even if $\\tau_\\ast$ is large. For a systematic fit, all available line profiles should be considered in order to adjust the parameters of the model consistently. This will be subject of a forthcoming paper. There are two aspects in the problem of fitting the observed line profiles. One is to reproduce the line shape, another is to provide for the observed line flux. Ignace \\& Gayley (\\cite{Rico02}) considered profile shapes for optically thick emission lines. They showed that these lines are blueshifted, and slightly narrower than optically thin lines. The lines have a universal shape which is nearly symmetric and insensitive to the level of continuous absorption. Nevertheless, the flux in the line scales inversely with optical depth of the cool wind, and therefore the problem of missing opacity is still unresolved in their approach. The same problem is highlighted in Kramer et al.\\ (\\cite{KCO03}), which up-to-day is the most consistent attempt to fit the observed lines. Our study shows that incorporation of wind fragmentation can resolve the controversy. A burning question is, why stars with existing high-resolution X-ray spectra show qualitatively different profile shapes, as listed in the introduction. At present we suggest that only $\\zeta$\\,Pup is a safe prototype for an isolated strong stellar wind. Indeed, speckle interferometry led to the discovery that $\\theta^1$\\,Ori\\,C and $\\zeta$\\,Ori\\,A, are binary systems (Hummel et al.\\ \\cite{Hummel00}, Weigelt et al.\\ \\cite{Weigelt99}), and $\\delta$\\,Ori\\,A is a well-known multiple system (see e.g. Miller et al.\\ \\cite{Milleretal02}). Moreover, $\\theta^1$\\,Ori\\,C has a strong magnetic field in excess of 1\\,kG (Donati et al.\\ \\cite{Don02}). The physical processes leading to the emission of X-rays can differ between objects. Nevertheless, we claim that the effects of wind fragmentation ought to be included in the modeling of emerging line profiles independent of the particular emission mechanism. \\begin{figure}[bth] \\centering \\epsfxsize=\\columnwidth \\mbox{\\epsffile{zoriNeX.ps}} \\caption{Profile of the Ne\\,{\\sc x} line from $\\zeta$\\,Orionis, as observed with the Chandra satellite (histogram). The smooth line is a tentative comparison with one of our calculations, taking the profile for $\\tau_\\ast = 10$ and $\\beta = 2$ from Fig.\\,\\ref{fig:bvthick}. The theoretical profile is scaled to the terminal wind velocity of $\\zeta$ Orionis (2100\\,km\\,s$^{-1}$) and convolved with the instrumental profile, a Gaussian of 0.023\\,\\AA\\ FWHM. The good fit demonstrates that our fragmented wind model can reproduce emergent line profiles of the observed shape.} \\label{fig:zeta_Ori} \\end{figure} The scenario chosen as a framework of the model presented here is that of Feldmeier et al.\\ \\cite{AF97}, where X-ray are produced in the collision of fast cloudlets with dense shells. The prominent feature of this hydrodynamic simulation is that the X-ray emitting plasma is always located at the starward face of the cool absorbing fragments. It was already shown in Paper\\,I, and is confirmed here, that this configuration leads to strong depletion of the central part of the line, an effect which is certainly not observed. The following conclusions can be drawn from our numerical modeling of the X-ray line emission from inhomogeneous, fragmented stellar wind: \\begin{enumerate} \\item Fragmentation drastically reduces the effective opacity of the wind. Therefore X-rays produced deep inside the wind can effectively escape, which would be totally absorbed in a homogeneous wind of the same mass-loss rate. \\item Absorption in a fragmented wind becomes effectively independent of the mass absorption coefficient and therefore of the wavelength, if the fragments are optically thick. \\item The line profiles from the fragmented wind model exhibit a variety of shapes, ranging from broad, blueshifted and flat-topped to narrow and nearly symmetric, and are thus of promising similarity to observations. \\item The possibility of a flat-topped, i.e.\\ almost symmetric blue part of the line profile is due to the model assumption that the fragments are flat, have small thickness, and are aligned perpendicular to the radial flow direction. \\item The effect of fragmentation is significant when the individual fragments are optically thick. Therefore the effect depends on the average number of fragments per radial direction. This parameter can be empirically restricted by a detailed line fit and has interesting implications for the theoretical understanding of the wind hydrodynamics. \\item The lateral size of the wind fragments has no influence on the emergent lines. \\item If the fragments are strictly confined to a pattern of radial cones while moving outwards, which ensures strict mass conservation in each radial direction, the resulting line profiles differ only little from the random fragment model where the absorbing shell fragments are randomly distributed in both radial and angular coordinates. \\item The numerical modeling featuring stochastically arranged emitting parcels of gas and absorbing fragments confirms the results of Paper\\,I, where analytical formulae have been derived from a statistical treatment. \\end{enumerate}" }, "0403/astro-ph0403531_arXiv.txt": { "abstract": "{Using spectra of normal emission line galaxies from the First Data Release of the Sloan Digital Sky Survey (SDSS) we have investigated the relations between the extinction $C({\\rm H}\\beta)$ as derived from the \\Ha/\\Hb\\ emission line ratio and various global parameters of the galaxies. Our main findings are that: 1) $C({\\rm H}\\beta)$ is linked with the galaxy spectral type and colour, decreasing from early- to late-type spirals. 2) $C({\\rm H}\\beta)$ increases with increasing metallicity. 3) $C({\\rm H}\\beta)$ is larger in galaxies with an older stellar population. 4) $C({\\rm H}\\beta)$ is larger for more luminous galaxies. 5) The extinction of the stellar light is correlated with both the extinction of the nebular light and the intrinsic galaxy colours. We propose phenomenological interpretations of our empirical results. We have also cross-correlated our sample of SDSS galaxies with the IRAS data base. Due to the lower redshift limit imposed to our sample and to the detection limit of IRAS, such a procedure selected only luminous infrared galaxies. We found that correlations that were shown by other authors to occur between optical and infrared properties of galaxies disappear when restricting the sample to luminous infrared galaxies. We also found that the optical properties of the luminous infrared galaxies in our SDSS sample are very similar to those of our entire sample of SDSS galaxies. This may be explained by the IRAS luminosity of the galaxies originating in the regions that formed massive stars less than 1\\,Myr ago, while the opacity of galaxies as derived from the \\Ha/\\Hb\\ emission line ratio is due to diffuse dust. We show some implications of our empirical results on the determination of global star formation rates and total stellar masses in normal galaxies. ", "introduction": "Accounting for the presence of dust is paramount for our understanding of the constitution and evolution of galaxies. Indeed, dust modifies the light we receive from galaxies by both dimming it and modifying its colour. The determination of the bolometric luminosity of galaxies, the description of their stellar populations using galaxy colours, the estimates of the star formation rates using observed fluxes in either the emission lines or ultraviolet continua, all depend on a correction for the effects of dust. Dust in itself is also an important constituent of galaxies, not so much by its mass but essentially by the effects it has on the thermal balance of the interstellar medium and on the formation of H$_2$ molecules, which has important consequences on the efficiency of star formation (e.g. Omukai \\cite{O00}, Hirashita \\& Ferrara \\cite{HF02}). Yet, the question of dust opacity of galaxies is still a subject of strong debate. Earlier, the presence of dust in galaxies was essentially inferred from its impact on the observed distribution of stellar light. The spectacular extinction in the edge-on Sombrero galaxy was attributed to dust located in its plane. The mottled aspect of the arms in face-on spiral galaxies was attributed to higher extinction in zones of star formation. Such observations developed the view that dust is a common constituent of spiral galaxies. In lenticular galaxies, dust lanes have been noticed (Sandage \\cite{S61}). Studies using various techniques have led to the conclusion that the average extinction decreases following the sequence late-type spirals -- early-type spirals -- lenticulars and ellipticals (see review by Calzetti \\cite{C01}). Inclination tests measure the dependence of galaxy disk surface brightness on inclination (e.g. Valentijn \\cite{V94} and references therein). This method is statistical by nature and strongly depends on sample selection, as discussed by Valentijn. Another method is to consider multiwavelength data of galaxies and solve simultaneously for the intrinsic colours of the stellar populations and for the reddening. The results of this method strongly rely on the adopted dust distribution models, as emphasized by Witt et al.\\,(\\cite{WTC92}), Bianchi et al.\\,(\\cite{BFG96}), and Witt \\& Gordon\\,(\\cite{WG00}). Infrared data from the IRAS survey provide detection of dust through its emission in the infrared. However, the interpretation of the infrared dust emission in terms of dust content is not straightforward (see e.g. Sauvage \\& Thuan \\cite{ST94}, Dale \\& Helou \\cite{DH02}). A different approach is to study the transparency of galaxies with respect to background light (see Keel \\& White \\cite{KW01} and references therein for methods using one nearby background galaxy, and Gonz{\\'a}lez et al. \\cite{GAD98} for methods using distant galaxies as background sources). Unfortunately, this method suffers from extremely low statistics. Another extinction indicator that can be used in galaxies presenting emission lines is the observed Balmer decrement. As is known, the intrinsic intensity ratios of hydrogen recombination lines in ionized nebulae have a negligible dependence on the emission conditions (Osterbrock \\cite{O89}) so that the observed values are a consequence of the dust extinction being different at the relevant wavelengths. This has for example been used in the influential work by Calzetti et al.\\,(\\cite{CKS94}) to determine an empirical extinction law for starburst galaxies. However, few studies have used this method to investigate the relation between extinction and overall galaxy type (e.g. Hubble type). In order to do this, one needs spectra of galaxies spanning the entire range of Hubble types. The atlas of Kennicutt (\\cite{K92}) provided such a data base for a limited sample of galaxies. Among these galaxies, 15 were considered by Sodr\\'e \\& Stasi\\'nska (\\cite{SS99}, hereinafter SS99) to be normal emission line galaxies, and these authors showed that the extinction at \\Hb\\, as measured by the \\Ha/\\Hb\\ ratio (in the remaining of the paper we will call it the \\emph{Balmer extinction}) decreases steadily from early-type to late-type spirals. Later, using the Nearby Field Galaxy Survey (NFGS) of Jansen et al. (\\cite{JFFC00a}, \\cite{JFFC00b}), which provided adequate data for about 100 galaxies, Stasi\\'nska \\& Sodr\\'e (\\cite{SS01}, hereinafter SS01) showed that redder galaxies have larger Balmer extinction. The Sloan Digital Sky Survey (SDSS), which aims at obtaining spectra of $10^{6}$ galaxies in the nearby Universe, provides a wonderful opportunity to study this question in more detail. The present paper makes use of the observations from the First Data Release (DR1; Abazajian et al. 2003, see also Stoughton et al. 2002) of the SDSS to study the relation between the Balmer extinction and other global properties of the galaxies. Section 2 explains the selection of the observational sample and quantities used in the analysis. Section 3 presents the relation between Balmer extinction and other galaxy parameters. Section 4 presents an interpretation of our results. In Section 5 we cross-correlate the galaxies of our sample with galaxies detected by IRAS, looking for additional clues on the origin of the observed extinction. Section 6 summarizes the main conclusions of this work and outlines some prospects. \\begin{figure} \\centerline{\\includegraphics[width=8cm]{stasinska_f1.ps}} \\caption{Examples of two fits using our code to measure the equivalent widths and fluxes of H$\\beta$ (top panels) and H$\\alpha$ (bottom panels). The left and right figures are for spectra with signal-to-noise ratio in $g$-band of 5.9 and 12.4, respectively. We also show the equivalent widths (EW) in \\AA\\ and signal-to-noise ratio (S/N) of the measured emission lines.} \\label{ews1} \\end{figure} ", "conclusions": "We have thus found (confirming the results of SS99 and SS01) that the Balmer extinction of galaxies decreases steadily from early- to late-type spirals (whether classified according to their spectral type or according to their colour). We have also shown that there is a direct dependence of Balmer extinction on metallicity, and also on the age of the stellar populations and on the total galaxy luminosity. This last statement is consistent with the finding of Wang \\& Heckman (\\cite{WH96}), based on relations using far-ultraviolet and far-infrared fluxes, that ``the optical depth of normal galactic disks increases with galaxy luminosity''. On the other hand, the dependence we find on galaxy type seems to contradict the general opinion summarized by Calzetti (\\cite{C01}) and the recent result by Kauffmann et al. (2003) who conclude that ``galaxies with the youngest stellar populations are the most attenuated by dust''. While previous studies were sometimes based on small samples, the result of Kauffmann et al. (2003) comes from a model fitting of the continua of $10^{5}$ galaxies from the SDSS. It should be noted that what they measure with this elaborate procedure is the attenuation of the \\emph{stellar} light. Our work concerns the extinction of \\emph{nebular} light and the determination of the Balmer extinction is straightforward. The interpretation in terms of global opacity of galaxies is however not simple: with \\Ha/\\Hb\\ we measure some sort of average opacity of the zones dimming the light from \\hii\\ regions (excluding the most opaque ones). Still, it is striking to see how well $C({\\rm H}\\beta)$ correlates with the galaxy spectral type. In this section, we present a possible phenomenological interpretation of our results, and show that our findings can actually be reconciled with the results of Kauffmann et al. (2003). \\begin{figure} \\centerline{\\includegraphics[width=5.cm]{stasinska_f10.ps}} \\caption{Histogram of the values of the Balmer extinction at \\Hb, $C(\\Hb)$ in our sample of galaxies. } \\label{fig8} \\end{figure} \\subsection{A phenomenological interpretation of the observed Balmer extinction trends in our sample} Let us write $C({\\rm H}\\beta)$ as the following product: \\begin{equation} C({\\rm H}\\beta) = n_{\\rm d} l \\sigma_{\\rm d} = \\frac{n_{\\rm d}}{n_{\\rm M}} \\frac{n_{\\rm M}}{n_{\\rm H}} \\frac{n_{\\rm H}}{n_{*}} n_{*} l \\sigma_{\\rm d} \\end{equation} where $n_{\\rm d}$ is average number of dust particles per unit volume in a galaxy, $n_{\\rm M}$ is average number of metallic atoms (in any form) per unit volume, $n_{\\rm H}$ is the average number of hydrogen particles per unit volume, $n_{*}$ is the average number of stars per unit volume, $l$ is the geometrical thickness of the region responsible for the optical extinction, and $\\sigma_{\\rm d}$ is the typical extinction cross section of dust grains. The common view that the opacity should increase from early- to late-types is comforted by such arguments that ``galaxies with young stars contain more gas and hence more dust than galaxies with old stellar populations'' (Kauffmann et al. \\cite{K03}). On the other hand, early-type galaxies are more metal-rich and on average more massive than late-type galaxies, so it is not necessarily surprising that the extinction is actually larger for late-type galaxies. In Sect. 3.2 we have given arguments to say that the Balmer extinction is also determined by the mean age of the stellar population. This could indicate that in late-type galaxies $n_{\\rm d}/n_{\\rm M}$ is larger than in early-type ones. Such a view seems indeed to be supported by recent models for the evolution of dust in galaxies (Hirashita \\cite{H99}) which take into account the processes of formation and destruction of dust (note however that there are presently many assumptions and uncertainties in such models, see e.g. Dwek \\cite{D98} or Edmunds \\cite{E01}). In these models, condensation in cool stellar winds from low-mass stars is an important source of dust production. We note that 75\\% of the galaxies of our sample have a measured $C({\\rm H}\\beta)$ between 0.3 and 0.9, as seen from the histogram shown in Fig. 10. On the other hand, judging from the metallicities derived for galaxies of luminosities similar to those of our sample (Zaritsky et al. \\cite{ZKR94}, Charlot et al. \\cite{CKL02}), the metallicity range in our sample is likely higher than just a factor of three. This suggests that factors other than metallicity act to reduce the observed range in $C({\\rm H}\\beta)$. Obviously, $n_{\\rm H}/n_{*}$ is a good candidate, since it decreases from late- to early-type galaxies (Roberts \\& Haynes \\cite{RH94}). On the other hand $ n_{*} l$, which can be assimilated to the stellar surface density, decreases from early- to late-types. The last factor in Eq. (3) is $\\sigma_{\\rm d}$, and one might expect some systematic effects if the grain size distribution depends on the processes for grain growth or destruction that could have different relative importances in galaxies of different types. In conclusion, dust extinction in galaxies involves many factors, and our finding that $C({\\rm H}\\beta)$ increases from late- to early-types can easily be accounted for, at least qualitatively, within our present-day understanding of galaxies and their constituents. \\begin{figure} \\centerline{\\includegraphics[width=5.13cm]{stasinska_f11.ps}} \\caption{The relation between \\EWHb and \\EWHa\\ in our sample of galaxies. The layout of the figure is the same as for Fig. 5. } \\label{fig9} \\end{figure} \\subsection{The extinction of stellar light versus the extinction of nebular light} As emphasized above, what we measure with $C({\\rm H}\\beta)$ is actually the reddening of the nebular light (called $C_{l}$ in SS01). On the other hand, what is determined by Kauffmann et al. (2003) is the reddening of the stellar light (called $C_{c}$ in SS01). We now examine the relation between these two quantities in our sample of galaxies. It had already been noted by SS99 and SS01, on much smaller samples of galaxies, that \\EWHa\\ and \\EWHb\\ correlate extremely well (see also Kennicutt 1992). This implies that the difference between the extinction of the stellar continuum and that of the nebular emission is strongly linked to the colours of the galaxies, being larger for redder galaxies. (see Eq. 5 in SS01). In Fig. 11 we show the values of EW(H$\\beta$) as a function of EW(H$\\alpha$) in our sample of galaxies. The correlation is extremely strong, with $\\tau_{K} = 0.782$ and $r_{S} = 0.931$. Assuming that the relation between these emission line equivalent widths is EW(H$\\beta$) = $A \\times$ EW(H$\\alpha$) (the same model as adopted in SS01), an ordinary least square bisector linear fitting (Isobe et al. 1990) gives $A = 0.185 \\pm 0.001$. Considering only objects for which EW(H$\\alpha$) $>$ 20 \\AA\\ the result is the same, $A = 0.185 \\pm 0.001$. These results may be compared with $A = 0.194 \\pm 0.011$ and $A = 0.245 \\pm 0.007$, presented in SS99 and SS01, respectively. Our value $A$ is consistent with that obtained by SS99 but not with that obtained by SS01. The cause of this discrepancy is not clear; it might be due to differences in the spectral resolution, because the spectral resolution of the SDSS spectra is almost half that of the NFGS spectra (Jansen et al. \\cite{JFFC00a}, \\cite{JFFC00b}) used by SS01. From Eq. (5) in SS01 and the value of $A$, we derive that \\begin{equation} C_{c} = C_{l} - 0.81 - 2.99 ~ {\\rm log}\\, \\frac {F_{c}^{o} ({\\rm H}\\alpha)}{F_{c}^{o}({\\rm H}\\beta)}, \\end{equation} where $F_{c}^{o}$(H$\\alpha$) and $F_{c}^{o}$(H$\\beta$) are the intrinsic (i.e. not affected by extinction) stellar fluxes in the continuum adjacent to \\Ha\\ and \\Hb\\ respectively. We can use the dust-free spectrophotometric models of Barbaro \\& Poggianti (1997) to relate the value of $F_{c}^{o}$(H$\\alpha$)/$F_{c}^{o}$(H$\\beta$) to the values of $D(4000)$ and \\EWHa. Using panel $j$ of our Fig. 6 we can relate \\EWHa\\ to log (\\Ha/\\Hb) (by taking the median value of \\Ha/\\Hb\\ for a given \\EWHa) and therefore to $C_{l}$. With the help of Eq. (4) we thus find that $C_{c}$ goes from about 0.02 for $D(4000) = 1.7$ to about 0.2 for $D(4000) = 1.3$. This goes in the same direction as the results of Kauffmann et al. (2003) who find that the median value of $A_{z}$, which is approximately equal to our $C_{c}$, roughly goes from 0.2 at $D(4000) = 1.7$ to 0.6 at $D(4000) = 1.3$. That the numbers are not exactly the same as the ones found by our analysis is not necessarily a worry given the dispersion in the observational points (both here and in the work of Kauffman et al. 2003) and given that the definition of the derived extinction is not exactly the same. As for $C_{l}$, it can be evaluated using Fig. 6c and Eq. 1. We find $C_{l}$ $\\simeq$ 0.8 for $D(4000) = 1.7$ and $C_{l}$ $\\simeq$ 0.3 for $D(4000) = 1.3$ We note that $C_{c}$ is smaller than $C_{l}$ at both extremes of the spectral type range, which can be interpreted as due to the fact that dust is more concentrated (and thus more opaque to radiation) in molecular clouds associated with \\hii\\ regions than in the diffuse interstellar medium. One may wonder why $C_{c}$ increases from early- to late-type spirals while $C_{l}$ decreases. The answer to this may be related to the fact that the stellar light from early-type galaxies is dominated by the bulge and to a specific distribution of dust resulting from dynamical effects. Advanced 3D-modelling of the star, dust and gas distribution in galaxies would be needed to test any interpretation of our empirical result. \\begin{figure*} \\centerline{\\includegraphics[width=10.9cm]{stasinska_f12.ps}} \\caption{Galaxies from our sample detected by IRAS (see Sect. 5). Plots of log (\\Ha/\\Hb) vs. log $L(\\rm FIR)$ (panel $a$) and log $L(\\Hb)$ (corrected for Balmer extinction) vs log $L(\\rm FIR)$ (panel b). } \\label{fig10} \\end{figure*} \\begin{figure*}[t] \\centerline{\\includegraphics[width=16cm]{stasinska_f13.ps}} \\caption{Same diagrams as in Fig. 6, with the galaxies detected by IRAS represented with large circles. } \\label{fig11} \\end{figure*} We have used the observations from the First Data Release of the SDSS to examine a sample of normal galaxies (as opposed to galaxies with an active nucleus) and to investigate how the Balmer extinction $C({\\rm H}\\beta)$ (i.e. the extinction at the wavelength of \\Hb\\ derived from the \\Ha/\\Hb\\ emission line ratios) relates with other global properties of the galaxies. Our selection criteria to build up the sample resulted in a data set of 9840 galaxies with adequate data. All these galaxies are at redhifts larger than 0.05 to avoid strong aperture effects. Our main findings are the following: \\begin{enumerate} \\item $C({\\rm H}\\beta)$ is linked with the galaxy spectral type and colour, decreasing from early- to late-type spirals. \\item $C({\\rm H}\\beta)$ increases with increasing metallicity \\item $C({\\rm H}\\beta)$ is, probably, also affected by the age of the stellar population, being larger in the case of older stellar populations. \\item $C({\\rm H}\\beta)$ depends on galaxy masses. \\item The extinction of the stellar light is correlated with both the extinction of the nebular light and the intrinsic galaxy colours, resulting in a trend with galaxy colour that may be opposite to the trend of $C({\\rm H}\\beta)$. \\end{enumerate} The present work thus confirms the conclusions of our previous studies (Sodr\\'e \\& Stasi\\'{n}ska \\cite{SS99} and Stasi\\'{n}ska \\& Sodr\\'e \\cite{SS01}), which were based on much smaller samples and used data with lower spectral resolution (Kewley et al. \\cite{KGJD02}, using the same sample as SS01, also found that early-types in that sample are more heavily reddened than late-types). Compared to our previous studies, the large number of galaxies in the SDSS sample allows us to investigate issues related to the inclination of galaxies. The fact that $C({\\rm H}\\beta)$ correlates so well with other properties of galaxies is remarkable, given that the extinction, especially in late-types, is known to be not uniform across the face of galaxies (e.g. Beckman et al. \\cite{BPK96}). We have cross-correlated our sample of SDSS galaxies with the IRAS data base in order to investigate any relationship between $C({\\rm H}\\beta)$ and total infrared luminosity of the galaxies. Due to the lower redshift limit imposed to our sample and to the detection limit of IRAS, such a procedure selected only luminous infrared galaxies. We found that correlations that were shown by other authors to exist between optical and infrared properties of galaxies disappear when restricting to luminous infrared galaxies. We also found that the optical properties of the luminous infrared galaxies in our SDSS sample are very similar to those of our entire sample of SDSS galaxies. We have proposed a phenomenological interpretation of our findings. We suggest that the main driver of the Balmer extinction of galaxies is their mass, combined with their metallicity and presence of old stellar populations. The infrared luminosity of the galaxies as determined by IRAS, which is attributed to radiation from hot stars reprocessed by dust grains, samples the regions with {\\it the most recent episodes} of star formation, and is not connected with the Balmer extinction. Obviously, detailed modelling of the spectral light from galaxies taking into account the effects of dust and using a complete code such as GRASIL (Silva et al. 1998, see also http://web.pd.astro.it/granato/grasil/grasil.html) is needed for a deeper understanding of the empirical relations we have found. This is not an easy task, however, since as noted by Witt at al. (1992) and Witt \\& Gordon (2000), equal amounts of dust in different configurations may produce very different reddening and attenuation effects. In any case, future models of the integrated light from galaxies including the effects of dust should also aim at reproducing the correlations we have shown. An important outcome of our study is to open the way for an improved correction for extinction in the determination of such parameters as the global star formation rate in galaxies or their total stellar masses. For normal galaxies, the global star formation rate can be obtained from the total \\Ha\\ luminosity corrected for extinction using the extinction derived from the \\Ha/\\Hb\\ emission line ratio (keeping in mind the reservations expressed e.g. by Hirashita et al. 2003). If observations do not allow one to determine the \\emph{Balmer extinction}, one can make use of e.g. the observed galaxy colour or the $D(4000)$ parameter to obtain an estimate of statistical value since we have shown that all these quantities are correlated. On the contrary, the total stellar mass can be estimated from the observed stellar fluxes of the galaxy after correcting for \\emph{stellar extinction}. This should be done with a proper model fitting of the observed continuum as in Kauffmann et al. (2003). However, the strong correlation that we have found empirically between stellar extinction, Balmer extinction and galaxy colours can provide a basis for a statistical method to determine the total masses of galaxies. These aspects will be developed in future work and should be important especially for the study of galaxies at intermediate and high redshifts." }, "0403/astro-ph0403641_arXiv.txt": { "abstract": "Evolution of a universe with homogeneous extra dimensions is studied with the benefit of a well-chosen parameter space that provides a systematic, useful, and convenient way for analysis. In this model we find a natural evolution pattern that entails not only stable extra dimensions in the radiation-dominated era, thereby preserving essential predictions in the standard cosmology, but also the present accelerating expansion while satisfying the limit on the variation of Newtonian gravitational constant. In this natural evolution pattern the extra dimensions tend to be stabilized automatically without resorting to artificial mechanisms in both the radiation-dominated and the matter-dominated era, as a wonderful feature for building models with extra dimensions. In addition, the naturalness of this evolution pattern that guarantees the late-time accelerating expansion of a matter-dominated universe presents a solution to the coincidence problem: why the accelerating phase starts at the present epoch. The feasibility of this evolution pattern for describing our universe is discussed. ", "introduction": "The observations by Supernova Cosmology Project and Supernova Search Team have suggested in 1998 that the expansion of the present universe is accelerating \\cite{Perlmutter:1999np,Riess:1998cb}. This conclusion is reinforced recently in 2003 by WMAP measurements \\cite{WMAP2003}. One general conclusion from these measurements and the CMB observations in recent years \\cite{Sievers:2002tq,Kuo:2002ua,WMAP2003} is that the universe has the critical density, consisting of $1/3$ of ordinary matter and $2/3$ of dark energy with a negative pressure \\cite{CMB&SN} (such that $p_{\\textsc{x}}/\\rho_{\\textsc{x}} < -0.78$ \\cite{WMAP2003}). Although this acceleration may be driven by the existence of a positive cosmological constant (vacuum energy) \\cite{Lambda models}, there remain other interpretations of the accelerating expansion, such as ``quintessence'' (a slowly evolving scalar field) \\cite{Caldwell:1998ii,ComplexQ} and the presence of extra dimensions \\cite{Gu:2001ni,Gu:2002mz}. The presence of extra dimensions is required in various theories beyond the standard model of particle physics, especially in the theories for unifying gravity and other forces, such as superstring theory. Extra dimensions should be ``hidden'' (or ``dark'') for consistency with observations. Various scenarios for ``hidden'' extra dimensions have been proposed, for example, a brane world with large compact extra dimensions in factorizable geometry proposed by Arkani-Hamed \\emph{et al.} \\cite{Arkani-Hamed} (see also \\cite{Antoniadis:1990ew}), and a brane world with noncompact extra dimensions in nonfactorizable geometry proposed by Randall and Sundrum \\cite{Randall&Sundrum}. In this paper we will employ the simplest scenario: small compact extra dimensions in factorizable geometry, as introduced in the Kaluza-Klein (KK) theories \\cite{Kaluza&Klein}. The possibility of generating the accelerating expansion of the present matter-dominated universe via the evolution of homogeneous and isotropic extra dimensions is first pointed out by Gu and Hwang \\cite{Gu:2001ni}. Furthermore, the general idea of unifying dark energy sources (i.e.\\ dark matter and dark energy) and dark geometry (e.g.\\ extra dimensions) into one has been sketched by Gu \\cite{Gu:2002mz}, who pointed out that dark geometry (instead of dark energy) can be an intriguing candidate for generating accelerating expansion. As indicated by Einstein's general relativity, mass (energy) and geometry are two faces of one single nature, therefore it is biased to consider only mass (energy) has dark part, while believing blindly that all geometry is totally ``visible'' to our poor eyes \\cite{Gu:2002mz}. In this simple scenario making use of a highly symmetric extra space, there are much fewer free parameters (only two additional degrees of freedom, the expansion rate and the curvature of the extra space, in addition to those in the standard cosmological models without dark energy) so that it is much easier to be ruled out, compared with the quintessence models, % by constraints from observations. In particular, an essential difficulty of this model \\cite{Cline:2002mw} stems from the constraint on the variation of the Newtonian gravitational constant, which will be produced along with the evolution of extra dimensions that is the key element for generating accelerating expansion. In this paper we will study a more general case in which the evolution of our universe in various eras, especially the present accelerating expansion era, is governed not only by the matter contents (excluding dark energy) therein, but also by the curvature of the ordinary 3-space and the evolution of extra dimensions. We will explore the feasibility of this model for generating accelerating expansion while satisfying observational constraints, especially the variation of the Newtonian gravitational constant. % Through these studies we will also present several nice features of this model, such as the automatic stabilization of extra dimensions and the solution to the cosmic coincidence problem --- why the energy densities of dark energy and dark matter are comparable now (i.e.\\ ``why now'' problem), or, more precisely (if dark energy is not a necessary ingredient for the accelerating expansion), why the accelerating phase of our universe starts at the present epoch. % ", "conclusions": "We have discussed the evolution of a universe under the influence of flat extra dimensions in various eras via a well-chosen parameter space that provides a systematic, useful, and convenient way for analysis. Since we assume the flatness (zero curvature) of the extra space, the only ingredient of extra dimensions that can affect the evolution of the universe is the evolution of their size. We have studied general features of the evolution, thereby exploring whether the picture is viable, subject to constraints from observations, for describing our world. Through the above studies we have found a natural evolution pattern from the blazing era to the matter-dominated era in this model. In the blazing era, the ordinary space and the extra space tend to synchronize their expansion rates, i.e., $v/u \\rightarrow 1$, and meanwhile the universe is decelerating, presuming that an extremely tiny curvature contribution $k_a/a^2 u^2$ has been guaranteed initially by inflation.\\footnote{This tendency corresponds to the fixed point at $(k_a/a^2 u^2 \\, , \\, v/u)=(0,1)$ in the parameter space for the blazing era, which is stable in the Y direction but unstable in the X direction.} This tendency provides a natural initial condition, $(k_a/a^2 u^2 \\, , \\, v/u)=(\\pm \\epsilon ,1)$, for the succeeding radiation-dominated era (where `$\\pm \\epsilon$' indicates a tiny curvature contribution). In the radiation-dominated era, the expansion rate of the extra space, $v$, tends to approach zero from $v/u \\simeq 1$ (as suggested above), meanwhile maintaining the decelerating phase.\\footnote{This tendency corresponds to the fixed point at $(k_a/a^2 u^2 \\, , \\, v/u)=(0,0)$ in the parameter space for the radiation-dominated era, which is stable in the Y direction but unstable in the X direction.} It is a good feature for building models with extra dimensions that the expansion rate of the extra space decreases to zero automatically, that is, extra dimensions are stabilized automatically in this era without resorting to any artificial mechanism. In particular, for small enough $v$ we can recover the standard cosmology (without extra dimensions) and the essential predictions therein (in particular, Big Bang Nucleosynthesis) in the radiation-dominated era. Therefore in our model with extra dimensions the predictions in the standard cosmology in the radiation-dominated era can be preserved in a natural way. In the matter-dominated era, a natural initial condition, $(k_a/a^2 u^2 \\, , \\, v/u)=(\\pm \\epsilon , 0)$, is provided by the preceding, radiation-dominated era. For the case of a negative curvature, i.e.\\ with the initial condition $(-\\epsilon , 0)$, the universe will eventually change its initial decelerating phase to the accelerating one and approach the attractor at $(k_a/a^2 u^2 \\, , \\, v/u)=(-1,0)$ with stable extra dimension and significant negative curvature contribution. As we have shown, this evolution pattern is marginally consistent with the limit on the variation of the Newtonian gravitational constant. The maximal acceleration associated with this evolution pattern deviates from the one required by SN Ia data together with CMB and LLS observations (within the framework of the standard cosmology) \\cite{Perlmutter:1999jt} with $2.5 \\sigma$ deviation. Thus this evolution pattern is still not ruled out by the present observations. We note that in this natural evolution pattern the expansion rate of extra dimensions will eventually approach zero, consequently extra dimensions being stabilized automatically without resorting to any artificial mechanism, in both the radiation-dominated and the matter-dominated era. This is the key ingredient for generating accelerating expansion of the present universe as well as satisfying the limit on the variation of the Newtonian gravitational constant (both caused by the evolution of extra dimensions), meanwhile preserving essential predictions in the standard cosmology for the radiation-dominated era. In particular, we emphasize that the naturalness of this evolution pattern, whose behavior is not sensitive to the initial conditions, indicates a solution to the cosmic coincidence problem (``why now'' problem) of dark energy % --- why dark energy starts dominating the universe now, or, more precisely, why the accelerating phase starts at the present epoch. % The existence of this natural evolution pattern implies that the late-time accelerating expansion of an open universe \\footnote{Actually this open universe is nearly flat in the earlier time.} in the matter-dominated era is guaranteed, accordingly the cosmic coincidence problem being solved, in our flat ED model with the help of inflation in the very early time. Nevertheless, type-(d) trajectories in general entail a too short age of the universe % and negative curvature in the ordinary space rather than almost zero curvature suggested by CMB data within the framework of standard cosmology. These features might eventually rule out this evolution pattern for describing our universe. Further detailed studies about the consistency with observations are in progress. Moreover, in order to complete the picture of the scenario for a universe with extra dimensions and find the best-fit trajectory therein, a more general model with nonzero curvature in the extra space is also under investigation." }, "0403/astro-ph0403194_arXiv.txt": { "abstract": "Searches for CO emission in high-redshift objects have traditionally suffered from the accuracy of optically-derived redshifts due to lack of bandwidth in correlators at radio observatories. This problem has motivated the creation of the new COBRA continuum correlator, with 4~GHz available bandwidth, at the Owens Valley Radio Observatory Millimeter Array. Presented here are the first scientific results from COBRA. We report detections of redshifted CO($J=3\\rightarrow2$) emission in the QSOs \\smm \\ and VCV J140955.5+562827, as well as a probable detection in \\rxj. At redshifts of $z=2.846$, $z=2.585$, and $z=2.796$, we find integrated CO flux densities of $5.4~\\jykms$, $2.4~\\jykms$, and $2.9~\\jykms$ for \\smm, VCV J140955.5+562827, and \\rxj, respectively, over linewidths of $\\Delta V_{FWHM} \\sim 350~\\kms$. These measurements, when corrected for gravitational lensing, correspond to molecular gas masses of order $M(\\hh) \\sim 10^{9.6-11.1}\\,\\msun$, and are consistent with previous CO observations of high-redshift QSOs. We also report $3\\sigma$ upper limits on CO(3$\\rightarrow$2) emission in the QSO \\lbqs \\ of $1.3~\\jykms$. We do not detect significant 3~mm continuum emission from any of the QSOs, with the exception of a tentative ($3\\sigma$) detection in \\rxj \\ of $S_{3mm}=0.92~\\mjybeam$. ", "introduction": "High-redshift QSOs have been shown to be often associated with dusty host galaxies (e.g.\\ Barvainis \\& Ivison 2002b), and as such offer much potential as powerful probes of the formation and evolution of massive galaxies in the universe. The masses of high-redshift host galaxies, as inferred by extrapolation from the locally measured correlation between central, supermassive QSO black hole masses and the velocity dispersion of stars in the host galaxy spheroids \\citep{gms2000,fm2000} are large. Thus, high-redshift QSOs are signposts to the position and redshift of massive galaxies in the early universe. Many high-redshift QSOs have been detected by recent millimeter and submillimeter surveys (e.g.\\ Carilli et al.\\ 2001; Isaak et al.\\ 2002; Barvainis \\& Ivison 2002b), implying that there is a link between QSOs and the massive, dusty, and gas-rich galaxy population discovered by deep submillimeter surveys (e.g.\\ Smail, Ivison, \\& Blain 1997; Cowie, Barger, \\& Kneib 2002). Most of these so-called submillimeter sources are too faint for optical redshifts to be determined, even with the largest telescopes, precluding deep searches for the large molecular gas reservoirs associated with young, massive galaxies. By studying large samples of dusty, high-$z$ QSOs and the gas and dust properties of their host systems, we can evaluate their overlap with the submillimeter galaxy population and investigate their properties and evolution. Recently, studies of CO emission in high-redshift objects have seen much success, with over 20 separate galaxies detected (e.g.\\ Barvainis, Alloin, \\& Bremer 2002a; Guilloteau et al.\\ 1999), the majority from galaxies with $S_{850\\micron}>10$~mJy (e.g.\\ Figure~1 in Isaak et al.\\ 2002). From these measurements it is possible to infer the total molecular gas mass in the high-$z$ host galaxies, and in turn establish the fraction of the total mass that is still in gaseous form -- a suggestive indicator of evolutionary state. CO measurements also provide long-sought evidence of previous metal enrichment, both in these objects and in the early universe, since carbon and oxygen are mainly produced by fusion reactions in stellar cores. These stars must also have produced copious amounts of other heavy elements. We therefore expect to see metal-rich gas ($Z \\sim Z_{\\sun}$) in massive starbursting systems very soon after the first starbursts (after $\\sim 200$~Myr). The spectacular detections of both CO \\citep{fabian_sdss,berCO}\\ and inferred presence of dust \\citep{berdust,sdssdust}\\ in SDSS~1148+5251 indicate that heavy-element enriched gas was already present in some objects at $z=6.4$, within 800~Myr of the Big Bang. Star formation must have already been an important process by $z \\sim 6.4$. The growing number of CO detections at lower, yet still cosmologically significant, redshifts, are very important, indicating how widespread star forming systems were at early epochs, and supporting the link between submillimeter galaxies and QSO hosts. Searches for high-redshift CO have been hampered by the limited spectrometer bandwidths (typically about 500-1500 $\\kms$) available to date on the current generation of millimeter telescopes and interferometers. The lack of bandwidth prevents easy search for CO emission because the optically-determined redshifts for QSOs, typically measured from broad emission lines, are known to be blueshifted from the host galaxy redshifts by as much as 1000-2000 $\\kms$ \\citep{tytlerfan}. As an example, the QSO APM 08279+5255 has an optical redshift of $z=3.87$ from \\citet{apmoptical}, determined through identification of its rest-frame UV emission lines; by contrast, the redshift of the CO line of its host galaxy, from \\citet{downes99}, is $z_{CO}=3.911$, a discrepancy of approximately 2500 $\\kms$ (blueshifted). At the Owens Valley Radio Observatory (OVRO), a new very broadband cross-correlator system, named COBRA, has been recently implemented to alleviate this problem. With a spectral coverage of 4 GHz, COBRA represents an eight-fold increase over the previous bandwidth at OVRO, and currently provides the largest available bandwidth for CO searches at millimeter wavelengths. This system is particularly advantageous for high-redshift spectral line searches when the redshift is poorly known \\emph{a priori}. Presented in Figure~1 is a sample COBRA spectrum, showing 3~GHz of the 4~GHz available bandwidth. As part of the correlator commissioning, we started a program to detect CO emission from high-redshift QSOs. To this end, we compiled a database of submillimeter-bright, $z>2$ QSOs from the literature from which to choose a sample of targets. To be included in the list, each object was required to meet the following criteria: (1) detection in 2 or more submillimeter/far-infrared (FIR) bands; (2) $T_{dust} \\leq 100$ K, as derived following \\citet{MinSED}, such that the rest-FIR spectral energy distribution includes a significant cold dust contribution; and (3) FIR luminosity greater than or equal to that of the prototypical ULIRG Arp 220 (i.e.\\ $L_{FIR}>10^{12.2}\\,\\lsun$). We chose a total of 4 QSOs from the total of 23 sources in this database to observe, selecting those with the greatest FIR luminosities but without a published CO detection: \\lbqs, \\smm, \\rxj, and VCV J140955.5+562827. In the sections that follow, we detail our observations and report the results of our efforts, which are part of an ongoing program of observations at OVRO. Throughout this paper, we assume an $\\Omega_{\\textrm{M}}=0.3$, $\\Omega_{\\Lambda}=0.7$ cosmology with $H_{0}=70~\\kms~\\mpc^{-1}$. We also provide a parallel analysis in parentheses in the Einstein-deSitter cosmology $\\Omega_{\\textrm{M}}=1$ and $\\Omega_{\\Lambda}=0$, with $H_{0}=75~\\kms~\\mpc^{-1}$. ", "conclusions": "Three more high-$z$ QSOs have been detected in CO emission, \\smm \\ at $z=2.846$, VCV J140955.5+562827 at $z=2.585$, and \\rxj \\ at $z=2.796$, using the new COBRA correlator at the Owens Valley Radio Observatory. All three submillimeter-bright QSOs possess large reservoirs of molecular gas, further linking submillimeter galaxies to their QSO contemporaries, and \\smm \\ may be one of the most massive CO systems known." }, "0403/hep-ph0403248_arXiv.txt": { "abstract": "\\widetext We propose a solution to the cosmological monopole problem: Primordial black holes, produced in the early universe, can accrete magnetic monopoles before the relics dominate the energy density of the universe. These small black holes quickly evaporate and thereby convert most of the monopole energy density into radiation. We estimate the range of parameters for which this solution is possible: under very conservative assumptions we find that the black hole mass must be less than $10^9$gm. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403311_arXiv.txt": { "abstract": "{ We investigated the main-line spectral evolution with shell thickness of oxygen rich AGB stars. The study is based on a sample of 30 sources distributed along the IRAS colour-colour diagram. The sources were chosen to trace the Miras with thick shells and the whole range of OH/IR stars. The Miras exhibit a 1665~MHz emission strength comparable to that at 1667~MHz. Even though the Miras of the study have quite thick shells, their spectral characteristics in both main lines attest to a strong heterogeneity in their OH shell with, in particular, the presence of significant turbulence and acceleration. The expansion velocity has been found to be about the same at 1665 and 1667~MHz, taking into account a possible velocity turbulence of 1-2~km~s$^{-1}$ at the location of the main-line maser emission. An increase in the intensity ratio 1667/1665 with shell thickness has been found. A plausible explanation for such a phenomenon is that competitive gain in favour of the 1667~MHz line increases when the shell is getting thicker. There is an evolution in the spectral profile shape with the appearance of a substantial inter-peak signal when the shell is getting thicker. Also, inter-peak components are found and can be as strong as the external standard peaks when the shell is very thick. This trend for an increase of the signal in between the two main peaks is thought to be the result of an increase of the saturation with shell thickness. All sources but two - a Mira and an OH/IR star from the lower part of the colour$-$colour diagram - are weakly polarized. The strong polarization observed for those two particular objects is thought to be the result of perturbations in their shells. ", "introduction": "\\label{section Intro.} The standard model proposed by Reid et~al. (1977) explains convincingly the double-peaked profile usually observed for Miras and OH/IR objects, particularly at 1612~MHz. It also explains the general features of the circumstellar envelopes. Nevertheless, many stars do not have such a simple OH maser profile. In particular, in the main lines, the peaks are often broad and consist of many components. To explain these more complex profiles new models have been suggested. Alcock \\& Ross (1986) have shown that it is possible to change the profile shape for a partially or fully saturated maser when the masering region becomes sufficiently thick. When that happens, the profiles consist of broader peaks and a signal is seen at the velocity of the star. This is in agreement with the spectral profiles observed for Miras and OH/IR objects, but the inferred internal and external radii disagree with the MERLIN and VLA interferometric observations. To overcome this discrepancy, Alcock \\& Ross suggest that the mass loss is not a continuous and uniform process but rather consists of ejection of blobs of material in random directions such that the symmetrical geometry is preserved. The model calculation results of Collison \\& Nedoluha (1995) agree with the conclusions by Alcock \\& Ross that the observed OH profile cannot be produced by the standard model description of a smooth, spherically symmetric, steady windlike mass loss. The assumption of a non-smooth shell is justified by the interferometric maps themself. Indeed, with the increase in interferometer resolution the more detailed maps revealed that the OH maser emission is located in clumps (Welty, Fix \\& Mutel, 1987; Chapman, Cohen \\& Saikia, 1991) and/or the shell shows evidence for deviation from spherical symmetry (Diamond et~al. 1985; Bowers, Johnston \\& de~Vegt 1989). Evidence for acceleration in the OH shell has been found for many stars (Chapman et~al. 1994; Etoka \\& Le~Squeren 1996, 1997; Szymczak et~al. 1998; Richards et~al. 1999). The importance of acceleration in the shape of the spectral profile is underlined by the work of Chapman \\& Cohen (1985). According to the value of the logarithmic velocity gradient $\\epsilon$, the shape of the observed line profile largely varies as follows~: (1) for $\\epsilon \\rightarrow 0$ the classical double-peaked profile is observed; (2) for $0 < \\epsilon < 1$ an inter-peak signal can be seen; (3) for $\\epsilon \\simeq 1$ a plateau-shaped profile is obtained and finally (4) for $\\epsilon > 1$ the profile is reduced to a single peak centred on the stellar velocity. Thus, many spectral profiles resembling those observed can be modelled by introducing acceleration in the OH shell. Based on the IRAS measurements at 12, 25 and 60~$\\mu$m, Olnon et~al. (1984) discovered the existence of a continuous sequence from Miras to OH/IR objects in the so-called colour-colour diagram which plots the colour indices $[60-25]$ vs. $[12-25]$. This sequence has been analysed in terms of shell thickness by van~der~Veen \\& Habing (1988) and van~der~Veen \\& Rugers (1989). These studies lead to a commonly accepted interpretation that it is an evolutionary track followed by intermediate-mass stars. In this scenario, Miras are the progenitors of OH/IR stars involving among other characteristics an increase of their mass-loss rate and expansion velocity with time. Nonetheless, the initial mass of the star could play a role in the setting up of that sequence. Based on an alternative model for the evolution of AGB stars proposed by Epchtein et~al. (1990), L\\'epine et~al. (1995) proposed that the IRAS colour-colour sequence could be a sequence of mass. The study presented here deals with the spectral profiles in both main lines of Miras and OH/IR stars distributed along the colour-colour diagram. The main aim is to determine whether there is a relation between the spectral profile and the thickness of the circumstellar shell. The next section presents the observations and selection criteria. Section~\\ref{section Results} and Appendices~\\ref{appendix A},~\\ref{appendix B} and \\ref{appendix C} present the results concerning the individual stars and a comparison with previous observations. The discussion is presented in Sect.~\\ref{section Discus.} and the final conclusions in Sect.~\\ref{section Conclu.}. ", "conclusions": "" }, "0403/astro-ph0403127_arXiv.txt": { "abstract": "We present 1.4\\,GHz \\hi\\ absorption line observations towards the starburst in NGC\\,2146, made with the VLA and MERLIN. The \\hi\\ gas has a rotating disk/ring structure with column densities between 6 and 18 $\\rm \\times 10^{21}\\:atoms\\, cm^{-2}$. The \\hi\\ absorption has a uniform spatial and velocity distribution, and does not reveal any anomalous material concentration or velocity in the central region of the galaxy which might indicate an encounter with another galaxy or a far--evolved merger. We conclude that the signs of an encounter causing the starburst should be searched for in the outer regions of the galaxy. ", "introduction": "Atomic hydrogen absorption measurements on (sub)arc\\-second scales allow us to investigate the distribution and dynamics of neutral gas in the inner regions of starburst galaxies and active galactic nuclei (AGN). While \\hi\\ emission measurements with current instruments are sensitivity--limited to a maximum resolution of a few arcseconds, absorption studies are limited only by the angular resolution of the telescope and the brightness distribution of the background radio continuum. We present \\hi\\ absorption measurements of the peculiar spiral galaxy NGC\\,2146. It is well established that NGC\\,2146 is undergoing a strong starburst, even stronger than that in M\\,82 (e.g.\\ {\\mbox{\\citealt{kronberg81}}}; \\citealt{tarchi00}; later TNG); however, the origin of this starburst is still unclear. A starburst is often triggered by an interaction with another galaxy, which perturbs the potential equilibrium of the gas, causing a flow of gas towards the center which results in an increase in density and thus fueling of the star--formation process. It therefore seems plausible that a starburst in a galaxy is triggered by an encounter if the galaxy belongs to a group which contains and still is connected by large \\hi\\ tails, for instance like M\\,81\\,--\\,M\\,82\\,--\\,NGC\\,3077 (\\citealt{yun94}, \\citealt{walter02}) and NGC\\,3627\\,--\\,NGC3628\\,--\\,NGC\\,3623 (the Leo triplet; \\citealt{zhang93}). Such an interaction has been searched for in NGC\\,2146 since the presence of a large \\hi\\ cloud is known since 1976. Using the NRAO 91--m telescope, \\citet{fisher76} detected a huge \\hi\\ cloud extending out to six Holmberg radii ($\\sim$\\,120\\,kpc) around NGC\\,2146. This cloud could be the consequence of a tidal interaction or of an explosion/ejection in the galaxy. However, there exists no kinematic evidence for the explosion hypothesis, and no companion has been found which may have interacted with NGC\\,2146. The higher resolution observations of the extended \\hi\\ around NGC\\,2146 with the WSRT \\citep{caspers86} and the VLA \\citep{tara96} resolved the cloud into a prominent tail, extending out to 90\\,kpc SE of the body of the galaxy, but no further evidence of an interaction was found at that time. In 1990, \\citet{hutch90} suggested instead that NGC\\,2146 appears to be in the final stage of a fairly gentle far--evolved merger, with the dominant galaxy (NGC\\,2146) now seen close to edge--on and the stripped companion being on a final plunge toward its center. The putative traces of this merger are, however, not particularly compelling. Evidence of a collision with another galaxy which did not remain embedded in NGC\\,2146 has been suggested as an alternative triggering mechanism by \\citet{young88}. They drew especially attention to the 10\\,kpc extended semi--arc of \\hii\\ regions (observed in H$\\alpha$ and [SII]), which is not coplanar with the rotating disk of the galaxy. A more recent analysis by \\citet{tara01} (later TPB) of the \\hi\\ stream suggests today a tidal interaction between NGC\\,2146 and a Low Surface Brightness (LSB) companion of which a remnant is apparently still seen as a 1.5$\\times$10$^{8}$\\,M$_{\\sun}$ \\hi\\ concentration in the southern tail. In order to study the kinematics and density of gas in the central region of NGC\\,2146 in the light of the proposed merger/encounter hypothesis, we have mapped the \\hi\\ absorption towards the nuclear radio continuum emission using the VLA (1\\farcs8 resolution = 130\\,pc) and MERLIN (0\\farcs2 resolution = 15\\,pc). At the distance of NGC\\,2146 (14.5\\,Mpc, \\citealt{benvenuti75}), 1$''$ is equivalent to 70\\,pc. ", "conclusions": " \\noindent{-- the \\hi\\ absorption is observed over the entire continuum radiation background which coincides with the starburst region;} \\noindent{-- the \\hi\\ gas towards the starburst region has a column density between 6 and 18 $\\rm \\times 10^{21}\\:atoms\\:cm^{-2}$;} \\noindent{-- the \\hi\\ velocity field has the structure of a rotating disk and ring, as often found in the inner region of similar starburst systems;} \\noindent{-- the \\hi\\ absorption has a uniform spatial and velocity distribution without any indication of an encounter with another galaxy or a far-evolved merger; there is no indication of a bar;} \\noindent{-- the absence of anomalous gas concentration or velocities and the agreement of the \\hi\\ absorption rotation curve with those of different constituents (CO and H$\\alpha$), suggest that the inner gas has had time to relax after the encounter; the signs of an encounter causing the starburst must be searched for in the outer regions of the galaxy.}" }, "0403/astro-ph0403257_arXiv.txt": { "abstract": "We show that orbit-superposition dynamical models (Schwarzschild's method) provide reliable estimates of nuclear black hole masses and errors when constructed from adequate orbit libraries and kinematic data. We thus rebut two recent papers that argue that BH masses obtained from this method are unreliable. These papers claim to demonstrate that the range of allowable BH masses derived from a given dataset is artificially too narrow as a result of an inadequate number of orbits in the library used to construct dynamical models. This is an elementary error that is easily avoided. We describe a method to estimate the number and nature of orbits needed for the library. We provide an example that shows that this prescription is adequate, in the sense that the range of allowable BH masses is not artificially narrowed by use of too few orbits. This is illustrated by showing that the $\\chi^2$ versus BH-mass curve does not change beyond a certain point as more orbits are added to the library. At that point, the phase-space coverage of the orbit library is good enough to estimate the BH mass, and the $\\chi^2$ profile provides a reliable estimate of its errors. A second point raised by critics is that kinematic data are generally obtained with insufficient spatial resolution (compared to the BH radius of influence) to obtain a reliable mass. We make the distinction between {\\em unreliable} determinations and {\\em imprecise} ones. We show that there are several different properties of a kinematic dataset that can lead to {\\em imprecise} BH determinations (insufficient resolution among them), but none of the attributes we have investigated leads to an unreliable determination. In short, the degree to which the BH radius of influence is resolved by spectroscopic observations is already reflected in the BH-mass error envelope, and is not a hidden source of error. The BH masses published by our group and the Leiden group are reliable. ", "introduction": "There are almost twenty detections of massive black holes (BHs) in galaxy centers that employ the technique of orbit superposition modeling \\citep{vdm98,crvdb99, cdzdm99, geb00a, geb03, capp1459, verolme02b}. The orbit superposition technique is based on a method originally invented by \\citet{sch79}, who noted that the time-averaged orbits in a stationary potential can be summed (at finite spatial resolution) to match the mass distribution that gives rise to the potential, thereby producing an equilibrium dynamical model. The Schwarzschild method was first used to analyze kinematic data for evidence of BHs in galaxy centers by Richstone \\& Tremaine (1985) for M87 and Dressler \\& Richstone (1988) for M31 and M32. Those early models were spherically symmetric. In modern implementations of this method, large sets of orbits are run in a specified axisymmetric potential (based on the starlight distribution and a central point mass) and then a non-negative linear superposition of orbits is found that best matches the kinematic and photometric observations. The usual goal of this process is the determination of only two quantities, the stellar mass-to-light ratio $\\Upsilon$ (assumed to be independent of position) and the BH mass $\\mbh$. \\citet{valluri04} (hereafter VME) and \\cite{cretton04} have challenged the reliability of BH mass determinations obtained via orbit superposition modeling. Cretton \\& Emsellem also argue that the uncertainties noted by VME can be dealt with via regularization (smoothing the distribution of orbit weights). VME raise a long list of problems and features of the orbit-superposition analysis technique. They focus on two main issues: orbit insufficiency and data insufficiency. First, they contend that the BH mass determination is artificially narrowed through the use of too small an orbit library and imply, but do not show, that we have made this mistake. In addition they suggest that the absence of a flat-bottomed $\\chi^2$ profile as a function of BH mass indicates that this problem has occurred. Second, they suggest that the BH mass $\\mbh$ can only be determined if the radius of influence $r_i = G\\mbh/\\sigma^2$ where the BH dominates the stellar dynamics is well-resolved spectroscopically (here $\\sigma$ is the line-of-sight velocity dispersion). In each of their main points they have fastened on one facet of rather complex problems. We show in \\S 2 that the use of an orbit library which is too small is an easily avoided elementary trap. In \\S 3 we refer the reader to other work in which we show that we have adequate resolution to determine BH masses reliably, and argue that the presence of extensive radial kinematic coverage mitigates the uncertainties associated with a ``not well-resolved'' radius of influence. We also show that the ``flat-bottomed'' behavior of $\\chi^2$ is a feature whose presence depends on several features of the quality and extent of the kinematic data, not just the size of the orbit library. The principal result is that a $\\chi^2$ profile constructed using our method does indicate the range of acceptable BH masses in a manner easily judged by the readers of our papers. ", "conclusions": "We have made a point of distinguishing between {\\em imprecise} (the error bars are large) measurements of the BH mass $\\mbh$ and {\\em unreliable} ones (the error bars are wrong). Figure 1 shows that increasing the number of orbits beyond our standard prescription does not broaden the range of BH masses allowed by our method. Contrary to the assertion of VME, we must be using enough orbits. The issue of data sufficiency is far more complex. In \\S 3 we illustrate (or cite) several ways that data insufficiency can lead to imprecise $\\mbh$ estimates. We also argue that these estimates are {\\em imprecise}, not unreliable. Probably the best evidence for this is the fact that our $\\mbh$ estimates obtained with genuine orbit-superposition models lie within their own errors of repeat studies with greatly improved resolution \\citep{geb03}\\footnote {We have not investigated filled kinematic maps in radius and angle, as might be obtained with integral-field devices like SAURON. Especially where these data imply triaxiality (thereby violating the basic assumption of our axisymmetric models), there may be unexpected changes in $\\mbh$. }. Flat-bottomed profiles---a special case of imprecision---can be produced by kinematic data with limited radial or angular coverage, by sparsely sampled data, or by insufficient resolution. A fourth way to produce a flat bottomed $\\chi^2$ profile is the use of {\\em noiseless} datasets constructed from two-integral models \\citep{cretton04}. This procedure creates models that match the data perfectly. Since axisymmetric models have more freedom than two-integral models, a range of three integral distribution functions will---by construction---provide an {\\em exact} match to the data. Since $\\chi^2$ is bounded by zero, a flat bottom in the $\\chi^2$ plot is unavoidable. \\citet{cretton04} argue that regularization (smoothing the distribution of orbit weights) reduces the uncertainty in $\\mbh$. Our models use a form of regularization (maximum entropy) but only as a numerical technique to accelerate convergence: we iterate our models while steadily reducing the weight of the entropy to the best-fit criterion until there is no further change in the best-fit model. \\citet{verolme02b} study models with and without regularization for M32 and find no difference in the best fit BH mass. This paper has examined the two most serious recent criticisms of BH mass estimates from orbit superposition. We conclude that the issues raised by these criticisms have not led us to report erroneous BH masses. There are a number of other tests of our procedure that also indicate that our method is reliable: \\begin{enumerate} \\item our method has been successfully blind-tested against (spherical) Jaffe models containing BHs; \\item our method yields the same masses as the Leiden code when applied to the same data (for NGC 821 and M32); \\item BHs weighed by this method display the same distribution of residuals from the mass-velocity dispersion relation as BHs weighed via gas dynamics \\citep{dorcarn}. \\end{enumerate} The tests described above are posted on the Nuker Team Website \\newline (http://www.noao.edu/noao/staff/lauer/nuker.html). The BH masses published by our group (\\citealt{geb00a,bow01,geb03,sio04}) are obtained using an adequate number of orbits. The plots of $\\chi^2(\\mbh)$ we publish provide the reader with an accurate assessment of the precision of the results. So far as we can tell, BH masses published from the Leiden group (\\citealt{vdm98,crvdb99,verolme02b,capp1459}) also should be reliable." }, "0403/astro-ph0403582_arXiv.txt": { "abstract": "We present three-dimensional Monte Carlo radiative transfer models of a very young ($<10^{5}$ years old) low mass ($50\\,\\mathrm{M_{\\sun}}$) stellar cluster containing 23 stars and 27 brown dwarfs. The models use the density and the stellar mass distributions from the large-scale smoothed particle hydrodynamics (SPH) simulation of the formation of a low-mass stellar cluster by Bate, Bonnell and Bromm. Using adaptive mesh refinement, the SPH density is mapped to the radiative transfer grid without loss of resolution. The temperature of the ISM and the circumstellar dust is computed using Lucy's Monte Carlo radiative equilibrium algorithm. Based on this temperature, we compute the spectral energy distributions of the whole cluster and the individual objects. We also compute simulated far-infrared Spitzer Space Telescope (\\emph{SST}) images (24, 70, and 160~$\\mathrm{\\mathrm{\\mu m}}$ bands) and construct colour-colour diagrams (near-infrared \\emph{HKL} and \\emph{SST} mid-infrared bands). The presence of accretion discs around the light sources influences the morphology of the dust temperature structure on a large scale (up to a several $10^{4}$~au). A considerable fraction of the interstellar dust is underheated compared to a model without the accretion discs because the radiation from the light sources is blocked/shadowed by the discs. The spectral energy distribution (SED) of the model cluster with accretion discs shows excess emission at $\\lambda=$~3--30~$\\mathrm{\\mu m}$ and $\\lambda>500\\,\\mathrm{\\mu m}$, compared to that without accretion discs. While the former is caused by the warm dust present in the discs, the latter is cause by the presence of the underheated (shadowed) dust. Our model with accretion discs around each object shows a similar distribution of spectral index (2.2--20~$\\mathrm{\\mathrm{\\mu m}}$) values (i.e. Class 0--III sources) as seen in the $\\rho$~Ophiuchus cloud. We confirm that the best diagnostics for identifying objects with accretion discs are mid-infrared ($\\lambda=$ 3--10~$\\mathrm{\\mu m}$) colours (e.g. \\emph{SST}~IRAC bands) rather than \\emph{HKL} colours. ", "introduction": "Systematic investigations of young stellar objects (YSOs) in a star-forming cloud including comparative studies of theoretical predictions and observations are important for understanding the stellar/substellar formation processes. Examples of well-studied star-forming clouds are the Orion Trapezium Cluster, NGC~2024, and the $\\rho$~Ophiuchus and Taurus-Auriga clouds. Recent observations of young clusters have been used to determine circumstellar disc frequencies, disc mass distributions, initial mass functions (IMFs) and evolutional stages of the objects in the clusters. It has been shown that \\emph{JHKL} and \\emph{HKL} colour-colour diagrams are particularly effective in identifying the presence of circumstellar discs (e.g. \\citealt{kenyon:1995}; \\citealt{McCaughrean:1995}; \\citealt{lada:2000}; \\citealt{haisch:2001}). The spectral indices \\citep{lada:1987} or the slopes of observed spectral energy distributions (SEDs) of YSOs from near- to mid-infrared wavelengths are used to classify SEDs, and the classification scheme is often related to the evolutionary stages of YSOs (e.g. \\citealt*{adam:1987}; \\citealt{myers:1987}). The distribution of circumstellar disk masses in young clusters can be measured from millimetre continuum emission (e.g. \\citealt{eisner:2003} for NGC~2024). Determining IMFs of young clusters requires evolutionary models (e.g. \\citealt{d-antona:1997}; \\citealt{baraffe:1998}) and infrared spectroscopy for spectral type identifications (e.g. \\citealt{luhman:1999} for the $\\rho$~Ophiuchus cloud). \\citet*{bate:2003a} presented results from a very large three-dimensional (3-D) smoothed particle hydrodynamics (SPH) simulation of the collapse and fragmentation of a 50\\,$\\mathrm{M_{\\sun}}$ turbulent molecular cloud to form a stellar cluster. The calculation resolved circumstellar discs down to $\\sim10$~au in radius and binary stars as close as 1~au. Although some observational predictions, such as the IMF and binary fraction, may be gleaned directly from a hydrodynamical simulation of stellar cluster formation, the principal observable characteristics (optical, near-IR, IR, and sub-millimetre images and spectra) require further detailed radiative transfer modelling. The density distribution of such hydrodynamical calculations is very complicated, and the corresponding radiative transfer must be also performed in full 3-D. There are two basic approaches to 3-D radiative transfer problems: \\,grid based methods (e.g. finite differencing, short- and long- characteristic methods), and particle (photon) based methods, i.e. Monte Carlo. Examples of the first kind are \\citet*{stenholm:1991}, \\citet{folini:2003} and \\citet*{steinacker:2002}. Those of the second kind include \\citet{witt:1996}, \\citet{pagani:1998}, \\citet*{wolf:1999}, \\citet{harries:2000} and \\citet{kurosawa:2001a}. The advantages of the second approach are, for example, the flexibility to treat a complex density distribution and a complex scattering function. Readers are referred to \\citet{steinacker:2003} and \\citet{pascucci:2003} for more extensive discussion on the advantages and disadvantages of these two different methods. The temperature of the interstellar and circumstellar dust in the cluster must be calculated in order to determine the source function of the dust emission. The radiative equilibrium temperature of the dust particles can be found using the Monte Carlo method (e.g. \\citealt*{lefevre:1982}; \\citealt{wolf:1999}; \\citealt{lucy:1999b}; and \\citealt{bjorkman:2001}). The technique used by \\citet{lucy:1999b} takes into account the fractional photon absorptions between two events of a `photon packet'; hence, it works well even in the limit of low opacity. \\citet{bjorkman:2001} used the immediate re-emission technique, in which radiative equilibrium is forced at each interaction with the dust. A photon is re-emitted immediately after an absorption event using a product of the dust opacity and the difference between the Planck function with a current temperature and that with a new temperature corresponding to the radiative equilibrium of the dust that absorbed the photon. Unlike the method of \\citet{lucy:1999b}, this does not require a temperature iteration if the opacity is independent of temperature. Alternatively, \\citet*{niccolini:2003} used the ray-splitting method showing its effectiveness for both low and high optical depth media. Here we aim to simultaneously resolve the dust on parsec and sub-stellar-radius spatial scales, whilst including multiple radiation sources. To overcome the resolution problem, we have implemented an adaptive mesh refinement (AMR) scheme in the grid production process of the TORUS radiative transfer \\citep{harries:2000}. See also \\citet{wolf:1999}, \\citet{kurosawa:2001a} and \\citet{steinacker:2002} for a similar gridding scheme in a radiative transfer problem. The method described by \\citet{lucy:1999b} is used to compute dust temperatures in our models. The objectives of this paper are: 1.\\,to compute observable quantities by solving the radiative transfer problem using a complex density distribution from the SPH calculation by \\citet{bate:2003a}; 2.\\,to analyse the predicted observational properties of the cluster generated in the simulation of \\citet{bate:2003a} at a distance of 140~pc which corresponds to the distance of nearby star-forming regions such as Taurus-Auriga and the $\\rho\\,$Ophiuchus cloud (e.g. \\citealt*{bertout:1999}). In Section~2, we describe the details of our models. The results of the model calculations are given in Section~3. The conclusions are summarised in Section~4. \\begin{figure} \\begin{center} \\includegraphics[% clip, scale=0.45]{fig01.eps} \\end{center} \\caption{Top: The scattering (dotted line), the absorption (dashed line), and the total (solid line) opacities of the grain model, described in Section~\\ref{sub:Dust-Model}, are shown as functions of wavelength. Bottom: The corresponding albedo of the grains is shown as a function of wavelength.} \\label{fig:albedo} \\end{figure} ", "conclusions": "} We have presented three-dimensional Monte Carlo radiative transfer models of a very young low-mass stellar cluster with multiple light sources (23 stars and 27 brown dwarfs). The density structure and the stellar distributions from the large-scale SPH simulation of \\citet{bate:2003a} were mapped onto our radiative transfer grid without loss of resolution using an AMR grid. The temperature of the ISM and the circumstellar dust was computed using the Monte Carlo radiative equilibrium method of \\citet{lucy:1999b}. The results have been used to compute the SEDs, the far-infrared \\emph{SST} (MIPS bands) images, and the colour-colour diagrams of this cluster. We find that the presence of circumstellar discs on scales less than 10~au (Model II) influences the morphology of the temperature structure of the cluster (Figure~\\ref{fig:temperature-map}), and can affect the temperature of the dust on large scales (up to a several $10^{4}$~au). The dust shadowed by accretion discs has a lower temperature than the model without the discs (Model I). The radiation coming from within a few au of the light sources is anisotropic/bipolar because of the circumstellar discs. The cluster SEDs (Figure~\\ref{fig:sed_compare}) from Models I and II both show peaks around $2\\,\\mathrm{\\mu m}$ and $200\\,\\mathrm{\\mu m}$. The first peak corresponds to that of stellar emission, and the latter corresponds to emission from ISM dust with $T\\approx20\\,\\mathrm{K}$. The excess emission of Model II between $\\lambda=$~3--30~$\\mathrm{\\mu m}$ is due to warm dust in the accretion discs. The excess emission in the far infrared ($\\lambda>500\\,\\mathrm{\\mu m}$) seen in Model II is caused by the colder ($T<20\\,\\mathrm{K}$) dust. Assuming the distance to the cluster to be 140~pc, we have constructed simulated images (Figure~\\ref{fig:mips-images-limit}) for a \\emph{SST} MIPS (24, 70, and 160~$\\mathrm{\\mu m}$) observation using the appropriate diffraction limits, the sensitivity limits and the angular sizes of pixels. The emission at 24~$\\mathrm{\\mu m}$ traces the locations of the stars and brown dwarfs very well. No clear distinction can be made between the 70~$\\mathrm{\\mu m}$ image from Model I and Model II. The major structures in the 160~$\\mathrm{\\mu m}$ images appear to be same in both Models, but the emission from the isolated small structure in the lower half of the images is much weaker in Model II. In the simulated 24~$\\mathrm{\\mu m}$ images, 8 out of 50 objects are detected above the flux limit for Model I, and 18 out of 50 objects are detected for Model II. We note that 15 of the 27 brown dwarfs in the cluster (Model II) have $K<20$ and would therefore detectable via deep imaging, while those that are too faint tend to be the youngest, most deeply embedded sources. Using the flux levels between 2.2~$\\mathrm{\\mu m}$ and 10~$\\mathrm{\\mu m}$, the spectral indices of each SEDs were computed. The objects were then classified according to the spectral index values and the ratio of $\\log\\lambda F_{\\lambda}$ values at $\\lambda=160\\,\\mathrm{\\mu m}$ (\\emph{SST} MIPS) and at $\\lambda=2.2\\,\\mathrm{\\mu m}$ ($K$ band). We have found that 54 per cent of the objects are classified as Class~0, none as Class~I, none as flat spectrum, 22 per cent as Class~II and 24 per cent as Class~III for Model I. For Model II, 56 per cent of objects are classified as Class~0, 18 per cent as Class~I, 8 per cent as flat spectrum, 12 per cent as Class~II, and 6 per cent as Class~III. We also found that the spectral index distribution of Model~II ( with $K<20$ objects) is very similar to that of the $\\rho$~Ophiuchus cloud observation by \\citet{greene:1994}. Even though our objects are $0.07\\,\\mathrm{Myr}$ old, there are a mixture of Class 0--III objects; hence, the class does not necessary relate to the ages of YSOs. According to the simulated \\emph{HKL} and mid-infrared (\\emph{SST} IRAC) colour-colour diagrams (Figures~\\ref{fig:color-color-ukirt} and \\ref{fig:color-color-irac}), the disc objects (Model II) tend to be slightly redder than discless objects (Model I) in the \\emph{H-K} vs \\emph{K-L} diagram, but the two populations are not well separated. This can be understood from the model cluster SEDs (Figure~\\ref{fig:sed_compare}) that show no significant difference in the flux levels between $\\lambda=$~1--3~$\\mathrm{\\mu m}$. On the other hand, the disc objects and discless objects are clearly separated in the colour-colour diagram of the \\emph{SST} IRAC bands (Figure~\\ref{fig:color-color-irac}). As expected, we find that longer wavelengths are more efficient in detecting circumstellar discs. For example mid-infrared ($\\lambda=$~3--10~$\\mathrm{\\mu m}$) colours (e.g. \\emph{SST}~IRAC bands) are superior to \\emph{HKL} colours. We find that the intrinsic colours of the disc objects (Figure~\\ref{fig:color-color-irac}) show a distribution very similar to that found by \\citet{meyer:1997}. The work presented in this paper can be extended to study how the observable quantities (e.g. colours of stars in a young cluster) evolves with time, by performing radiative transfer model calculations with density structures from a hydrodynamics calculation at different times (ages). Growth of dust grain sizes in accretion discs may become important in predicting how the colours of objects evolve (e.g. see \\citealt{hansen:1974}; \\citealt{wood:2002b}). One of the obvious next steps in modelling star formation is to combine SPH and radiative-transfer in order to more accurately predict the temperature of the dust. It is interesting to note that the computational expense of the radiative equilibrium calculation performed here on one SPH time-slice (2400 CPU hours on UKAFF) is a significant fraction of that required to perform the complete SPH simulation (95\\,000 CPU hours on the same machine) which involves may thousands of time steps. Although it is not immediately clear how often the temperature would have to be computed during a hydrodynamic simulation (the radiative-transfer calculation would have its own Courant condition), it is apparent that a much more efficient method for calculating the radiation transfer is required before a combined SPH/radiative-transfer simulation becomes tractable." }, "0403/astro-ph0403541_arXiv.txt": { "abstract": "Accreting black holes are believed to emit X-rays which then mediate information about strong gravity in the vicinity of the emission region. We report on a set of new routines for the {\\sc{}xspec} package for analysing X-ray spectra of black-hole accretion disks. The new computational tool significantly extends the capabilities of the currently available fitting procedures that include the effects of strong gravity, and allows one to systematically explore the constraints on more model parameters than previously possible (for example black-hole angular momentum). Moreover, axial symmetry of the disk intrinsic emissivity is not assumed, although it can be imposed to speed up the computations. The new routines can be used also as a stand-alone and flexible code with the capability of handling time-resolved spectra in the regime of strong gravity. We have used the new code to analyse the mean X-ray spectrum from the long {\\it{}XMM--Newton} 2001 campaign of the Seyfert~1 galaxy MCG--6-30-15. Consistent with previous findings, we obtained a good fit to the broad Fe~K line profile for a radial line intrinsic emissivity law in the disk which is not a simple power law, and for near maximal value of black hole angular momentum. However, equally good fits can be obtained also for small values of the black hole angular momentum. The code has been developed with the aim of allowing precise modelling of relativistic effects. Although we find that current data cannot constrain the parameters of black-hole/accretion disk system well, the code allows, for a given source or situation, detailed investigations of what features of the data future studies should be focused on in order to achieve the goal of uniquely isolating the parameters of such systems. ", "introduction": "There is now strong evidence that the \\fekalfa line emission in some active galactic nuclei (AGN) and some Galactic X-ray binary black-hole candidates (BHC) originates, at least in part, from an accretion disk in a strong gravitational field. A lively debate is aimed at addressing the question of what the spectral line profiles and the associated continuum can tell us about the central black-hole, and whether they can be used to constrain parameters of the accretion disk in a nearby zone, about ten gravitational radii or less from the center. For a recent review for AGNs, see Fabian et al.\\ (2000), Reynolds \\& Nowak (2003), and references cited therein. For BHCs, see Miller et al.\\ (2002a), McClintock \\& Remillard (2003) and references therein. In several sources there is indication of \\fekalfa line emission from within the last stable orbit of a Schwarzschild black hole (e.g.\\ in the Seyfert galaxy MCG--6-30-15; see Iwasawa et al.\\ 1996; Fabian et al.\\ 2000; Wilms et al.\\ 2001; Martocchia, Matt \\& Karas 2002a) while in other cases the emission appears to arise farther from the black hole (e.g.\\ in the microquasar GRS 1915+105, see Martocchia et al.\\ 2002b; for AGNs, see Yaqoob \\& Padmanabhan 2004 and references therein). Often, the results from X-ray line spectroscopy are inconclusive, especially in the case of low spectral resolution data. For example, a spinning black hole is allowed but not required by the line model of the microquasar V4641 Sgr (Miller et al.\\ 2002b). The debate still remains open, but there are good prospects for future X-ray astronomy missions to be able to use the \\fekalfa line to probe the space-time in the vicinity of a black hole, and in particular to measure the angular momentum, or the spin, associated with the metric. One may also be able to study the `plunge region' (about which very little is known), between the event horizon and the last stable orbit, and to determine if any appreciable contribution to the \\fekalfa line emission originates from there (Reynolds \\& Begelman 1997; Krolik \\& Hawley 2002). In addition to the Fe~K lines, there is some evidence for relativistic {\\it soft X-ray} emission lines due to the Ly$\\alpha$ transitions of oxygen, nitrogen, and carbon (e.g.\\ Mason et al.\\ 2003), although the observational support for this interpretation is still controversial (e.g.\\ Lee et al.\\ 2001). With the greatly enhanced spectral resolution and throughput of future X-ray astronomy missions, the need arises for realistic theoretical models of the disk emission and computational tools that are powerful enough to deal with complex models and to allow actual fitting of theoretical models to observational data. It is worth noting that some of the current data have been used to address the issue of distinguishing between different space-time metrics around a black hole, however, the current models available for fitting X-ray data are subject to various restrictions which we shall elaborate on in the present paper. In this paper we describe a generalised scheme and a code which can be used with the standard X-ray spectral fitting package, {\\sc{}xspec} (Arnaud 1996). We have in mind general relativity models for black-hole accretion disks. Apart from a better numerical resolution, the principal innovations compared to currently available schemes are that the new model allows one to (i)~fit for the black-hole spin, (ii)~study the emission from the plunge region, and (iii)~specify a more general form of emissivity as a function of polar coordinates in the disk plane (both for the line and for the continuum). Furthermore, it is also possible to (iv)~study time variability of the observed signal and (v)~compute Stokes parameters of a polarized signal. Items (i)--(iii) are immediately applicable to current data and modelling, while the last two mentioned features are still mainly of theoretical interest at present. Time-resolved analysis and polarimetry of accretion disks are directed towards future applications when the necessary resolution and the ability to do polarimetry are available in X-rays. Thus our code has the advantage that it can be used with time-resolved data for reverberation studies of relativistic accretion disks (Stella 1990; Reynolds et al.\\ 1999; Ruszkowski 2000; Goyder \\& Lasenby 2004). Also polarimetric analysis can be performed, and this will be extremely useful because it can add very specific information on strong-gravitational field effects (Connors, Stark \\& Piran 1980; Matt, Fabian \\& Ross 1993; Bao et al.\\ 1997). Theoretical spectra with temporal and polarimetric information can be analysed with the current version of our code and such analysis should provide tighter constrains on future models than is currently possible. As mentioned above, Fe~K lines have been reported in the X-ray spectra of numerous AGNs and BHCs. These sources frequently exhibit remarkable variability patterns which are still difficult to understand. Among these puzzling objects, the Seyfert~1 galaxy MCG--6-30-15 is radically distinctive with its broad and skewed \\fekalfa feature that persists in observations in spite of substantial variablity of the continuum. Interpretation of the Fe~K line in terms of reflection from a relativistic black-hole disk has been found to be rather robust, but it has not been possible to fully explore the model parameter space. In particular, constraints on $a$ have only been obtained by fitting `inverted photon data' which is made from the real data in a model-dependent way that already assumes a certain value of $a$. The puzzling stability of this line has been attributed, amongst other things, to general relativity effects very close to the black hole (Miniutti et al.\\ 2003; Fabian \\& Vaughan 2003). It is thus interesting to know if and how a generalisation of the standard disk-line scheme together with accurate computation of general relativistic effects can add new pieces of information or whether a more substantial modification of the whole picture is needed (for further discussion and references, see e.g.\\ Weaver \\& Yaqoob 1998; Krolik 1999; Hartnoll \\& Blackman 2001; Dumont et al.\\ 2002). Different kinds of such generalisations have been proposed. Previous papers indicated that details of the model do matter. However, a strong gravitational field has almost invariably been required. In the present paper, we adopt the assumption that the line and continuum emission are produced by an irradiated disk near a rotating black hole, we relax some of the previous assumptions, and we search for best-fitting parameters of the model. In the \\S~2 we describe the main features of the new computational tool. We list several variants of the code which address different problems of fitting X-ray spectra of black hole plus accretion disk systems. In \\S~3 we use the code to analyse time-averaged data for MCG--6-30-15 {\\it{}XMM--Newton} CCD data as detailed in Fabian et al.\\ 2002). We point out that more complex models may not require a large value of the black-hole angular momentum, although the current data are consistent with a maximally rotating black hole. Finally, in \\S~4, we summarise our results and state our conclusions. \\begin{table*}[tbh] \\caption{Basic features of the new model in comparison with other black hole disk-line models.} \\begin{center} \\begin{small} \\tabletypesize{\\scriptsize} \\begin{tabular}{lccccl} \\tableline \\multicolumn{1}{c}{Model} & \\multicolumn{4}{c}{Effects that are taken into account\\rule[-1.5ex]{0mm}{4.5ex}} & \\multicolumn{1}{c}{Reference} \\\\ \\cline{2-5} & \\rule[-3.5ex]{0mm}{8ex}\\parbox{29mm}{\\footnotesize{}Energy shift of photons/Lensing effect} & \\parbox{18mm}{\\footnotesize{}Black hole angular momentum} & \\parbox{18mm}{\\footnotesize{}Axisymmetry is assumed} & \\parbox{20mm}{\\footnotesize{}Steady source is assumed} & \\\\ \\tableline {\\sf{}diskline} \\rule{0mm}{3ex} & yes/no~ & $0$~ & yes & yes & Fabian et al.\\ (1989) \\\\ {\\sf{}laor} & yes/yes & $0.998$~ & yes & yes & Laor (1991) \\\\ {\\sf{}kerrspec} & yes/yes & $\\langle0,1\\rangle^{\\dag}$ & no~ & yes & Martocchia et al.\\ (2000) \\\\ {\\sf{}ky} & yes/yes & $\\langle0,1\\rangle$ & no$^{\\ddag}$ & no & This paper \\\\ \\tableline \\end{tabular} \\end{small} \\vspace*{2mm}\\par{} {\\parbox{0.9\\textwidth}{\\footnotesize{}$^{\\dag}$~The value of dimension-less $a$ parameter is kept frozen.\\\\ $^{\\ddag}$~A one-dimensional version is available for the case of an axisymmetric disk. In this axisymmetric mode, {\\sf{}ky} still allows $a$ and other relevant parameters to be fitted (in which case the computational speed of {\\sf{}ky} is then comparable to {\\sf{}laor}). The results can be more accurate than those obtained with other routines because of the ability to tune the grid resolution.}} \\label{tab:models} \\end{center} \\end{table*} \\begin{table}[!tbh] \\caption{Basic versions of the model.} \\begin{center} \\begin{footnotesize} \\tabletypesize{\\footnotesize} \\begin{tabular}{l@{~}p{22mm}@{~}p{36mm}} \\tableline Name \\rule[-1.5ex]{0mm}{4.5ex} & Type$^{\\dag}$~ & Usage \\\\ \\tableline {\\sf{}KYGline} & additive \\rule{0mm}{3ex} & Relativistic spectral line from a black hole disk. \\\\ {\\sf{}KYHrefl} & additive & Compton reflection with an incident power-law (or a broken power-law) continuum. \\\\ {\\sf{}KYLcr} & additive & Lamp-post Compton reflection model. \\\\ {\\sf{}KYSpot} & To be used as an independent code outside {\\sc{}xspec} & Time-dependent spectrum of a pattern co-orbiting with the disk. \\\\ {\\sf{}KYConv} & convolution & Convolution of the relativistic kernel with intrinsic emissivity across the disk. \\\\ \\tableline \\end{tabular} \\end{footnotesize} \\vspace*{2mm}\\par{} {\\vspace*{-1mm}\\parbox{0.45\\textwidth}{\\footnotesize{}$^{\\dag}$~Different model types correspond to {\\sc{}xspec} syntax and are defined by the way they act in the overall model and form the final spectrum. According to the usual convention in {\\sc{}xspec}, additive models represent individual {\\it emission} spectral components which may originate e.g.\\ in different regions of the source. Additive models are simply superposed in the total signal. Multiplicative components (e.g.\\ {\\sf{}hrefl}, discussed in the Appendix~\\ref{appb}) multiply the current model by an energy-dependent factor. Convolution models modify the model in a more non-trivial manner. See Arnaud (1996) for details.}} \\label{tab:versions} \\end{center} \\end{table} ", "conclusions": "We have presented an extended computational scheme which can be conveniently employed to examine predicted X-ray spectral features from black-hole accretion disks. As mentioned above, a one-dimensional version of the code basically reproduces previous methods, with the addition of the option to vary more parameters (e.g.\\ $a$). The current two-dimensional version also accomodates non-axisymmetric and time-evolving emissivity from a geometrically thin disk. The new tool facilitates accurate comparisons between model spectra and actual data. It provides also polarimetric information. One can even substitute other tables for the Kerr metric tables to explore different specetime. However, we defer detailed discussion of these latter capabilities to a future paper. Furthermore, work is in progress to include realistic emissivities relevant to different physical situations, in order to allow for the intrinsically three-dimensional geometry of the sources (e.g.\\ an optically-thin corona above the disk), and to parallelize the code. Recently, several research groups have embarked on projects to compare theoretical predictions of intrinsic emissivity of accretion disks with observational data obtained by X-ray satellites. This approach offers a fascinating opportunity to explore processes near black-hole sources, both AGNs and BHCs, and to estimate physical parameters of black holes themselves. In order to limit the number of unknown model parameters and to alleviate the computational burden of the fit procedure, simplifications on the model emissivity (for example, stationarity and/or axial symmetry) have been often imposed.\\footnote{Traditionally, Occam's razor is invoked to justify simplifying assumptions. However, this way of reasoning is not sufficiently substantiated here in view of the fact that accretion flows are normally found to be very turbulent and fluctuating rather than smooth and steady. Put in other words, the assumption of an axisymmetric and stationary disk may provide a successful fit to the time averaged spectrum, but it is not satisfactory because we know that more complicated models {\\it{}must} be involved on physical grounds. (``Things should be made as simple as possible---but not too simple\\ldots'')} Naturally, fitting data with complex models is computationally more demanding. To summarize, we have illustrated the code by following common practice and employing minimisation of $\\chi^2$ to find the best-fitting parameter values for the highest signal-to-noise Fe K line profile available to date---from the {\\it{}XMM--Newton} campaign of the Seyfert~1 galaxy MCG--6-30-15. The new model can vary more of the parameters than previous variants of the disk-line scheme allowed, which seems to be necessary in view of the fact that simple prescriptions for intrinsic emisivity (axisymmetric, radially monotonic, etc.) are not adequate to reflect realistic simulations of accretion flows. It is without much surprise that a complex model exhibits an intricate $\\chi^2$ space in which ambiguities arise with multiple islands of acceptable parameters. The mean {\\it{}XMM--Newton} spectrum of MCG--6-30-15 analyzed here justifies rejection of the often-used plain power-law emissivity that depends solely on radius from the black hole. However, we cannot arbitrate the issue of non-monotonic emissivity versus a non-axisymmetric one. Due to the inherent degeneracy in time-averaged spectra, both alternatives can reproduce these data with reasonable parameters. In other words, $a$ of the black hole cannot be unambiguously determined if complex (realistic) emissivities are adopted. It is clear that the relativistic line profiles affected by strong gravity are such that it is very difficult to recover unambiguous information about the key parameters, such as black-hole spin, disk inclination angle, and inner disk radius. This is because even with increased throughput, energy resolution and signal-to-noise, the smooth, featureless tails of the time-averaged line profiles do not retain sufficient information to break the degeneracies between the dependences on the black-hole, accretion disk, and line emissivity parameters. Higly specific spectral shapes could actually be provided by contributions to the line profile from higher-order images, whose form inherits unique information about space-time of the black hole. It may appear rather unlikely that higher-order images could be discovered in spectra of AGNs, because this would require to see an unobscured inner disk at large (almost edge-on) inclination. As the data improve, the important measurable quantities that will constrain parameters when compared with the models are: (i)~the slope of the red wing, (ii)~the energy of the cut-off of the blue side of the line profile, and (iii)~the detailed shape of the blue peak (see Figs.\\ \\ref{fig:example0}--\\ref{fig:example3} and further discussion in the Appendix). The extreme part of the red wing will not be as useful because it will always be difficult to tell where the line emission joins the continuum no matter how good the data are, and also because there is often no prominent sharp feature in the red wing since the red Doppler peak is too smeared out when the line emission extends too close to the black hole. In addition, time variability will be extremely important. No matter how, or if, the line emission responds to the continuum, we know that the disk inclination and black-hole spin very likely do not change during different snapshots. Therefore any differences in the snapshot line profiles must be due only to changes in the line emissivity, in terms of its radial and azimuthal distribution and/or its radial extent ($r_{\\rm{}in}$ and $r_{\\rm{}out}$). Combining these approaches and collecting enough snapshots may provide sufficient information to uniquely determine one of the parameters, such as black-hole spin." }, "0403/astro-ph0403588_arXiv.txt": { "abstract": "We present a comprehensive space--based study of ten X--ray luminous galaxy clusters ($L_X{\\ge}8{\\times}10^{44}{\\rm erg \\,s^{-1}}$[0.1--2.4\\,keV]) at $z{=}0.2$. \\emph{Hubble Space Telescope (HST)} observations reveal numerous gravitationally--lensed arcs for which we present four new spectroscopic redshifts, bringing the total to thirteen confirmed arcs in this cluster sample. The confirmed arcs reside in just half of the clusters; we thus obtain a firm lower limit on the fraction of clusters with a central projected mass density exceeding the critical density required for strong--lensing of $50\\%$. We combine the multiple--image systems with the weakly--sheared background galaxies to model the total mass distribution in the cluster cores ($R{\\le}500{\\rm kpc}$). These models are complemented by high--resolution X--ray data from \\emph{Chandra} and used to develop quantitative criteria to classify the clusters as relaxed or unrelaxed. Formally, $(30{\\pm}20)\\%$ of the clusters form a homogeneous sub--sample of relaxed clusters; the remaining $(70{\\pm}20)\\%$ are unrelaxed and are a much more diverse population. Most of the clusters therefore appear to be experiencing a cluster--cluster merger, or relaxing after such an event. We also study the normalization and scatter of scaling relations between cluster mass, luminosity and temperature. The scatter in these relations is dominated by the unrelaxed clusters and is typically ${\\sigma}{\\simeq}0.4$. Most notably, we detect 2--3\\,times more scatter in the mass--temperature relation than theoretical simulations and models predict. The observed scatter is also asymmetric -- the unrelaxed systems are systematically 40\\% hotter than the relaxed clusters at 2.5--${\\sigma}$ significance. This structural segregation should be a major concern for experiments designed to constrain cosmological parameters using galaxy clusters. Overall our results are consistent with a scenario of cluster--cluster merger induced boosts to cluster X--ray luminosities and temperatures. ", "introduction": "\\setcounter{footnote}{7} \\normalsize Massive galaxy clusters are the largest collapsed structures in the Universe ($M_{\\rm virial}{\\simeq}10^{15}{\\rm M_\\odot}$), containing vast quantities of the putative dark matter (DM), hot intracluster gas (${\\tx}{\\simeq}7{\\rm keV}$), and galaxies ($n_{\\rm gal}{\\sim}10^3$). These rare systems stand at the nodes of the ``cosmic web'' as defined by the large--scale filamentary structure seen in both galaxy redshift surveys (e.g.\\ de~Lapparent, Geller \\& Huchra 1986; Shectman et al.\\ 1996; Vettolani et al.\\ 1997; Peacock et al.\\ 2001; Zehavi et al.\\ 2002) and numerical simulations of structure formation (e.g.\\ Bond et al.\\ 1996; Yoshida et al.\\ 2001; Evrard et al.\\ 2002). Clusters are inferred to assemble by accreting matter along the filamentary axes, slowly ($t_{\\rm crossing}{\\sim}2$--3\\,Gyr) ingesting DM, gas and stars into their deep gravitational potential wells. Clusters have long been recognized as cosmological probes. For example, the evolution of cluster substructure with look--back--time is in principal a powerful diagnostic of the cosmological parameters (Gunn \\& Gott 1972; Peebles 1980; Richstone, Loeb \\& Turner 1992; Evrard et al.\\ 1993). A complementary probe is to constrain the matter density of the universe and the normalization of the matter power spectrum using the cluster mass function. However, it is currently not possible to measure the cluster mass function directly. More easily accessible surrogates such as the X--ray luminosity and temperature functions are therefore used in combination with scaling relations between the relevant quantities (e.g.\\ Eke et al.\\ 1996; Reiprich \\& Boehringer 2002; Viana, Nichol \\& Liddle 2002; Allen et al.\\ 2003). A critical component of such analyses is the precision to which the scaling relations are known. Samples of X--ray selected clusters are now of a sufficient size that systematic uncertainties may be comparable with the statistical uncertainties, and therefore deserve careful analysis before robust cosmological conclusions may be drawn (e.g.\\ Smith et al.\\ 2003). Measurements of the Sunyaev--Zeldovich effect (SZE) are also emerging as a powerful cosmological tool (Carlstrom, Holder \\& Reese 2002). Cosmological SZE surveys will rely on the cluster mass--temperature relationship in a similar manner to cosmological X--ray surveys. Such experiments may therefore also be compromised if astrophysical systematics are identified and carefully eliminated from the analysis (e.g.\\ Majumdar \\& Mohr 2003). Detailed study of the assembly and relaxation histories of clusters, and their global scaling relations as a function of redshift are therefore vitally important. To advance our understanding of the assembly, relaxation and thermodynamics of massive galaxy clusters requires information about the spatial distribution of DM, hot gas and galaxies in clusters. Several baryonic mass tracers are available, for example X--ray emission from the intracluster medium (hereafter ICM -- e.g.\\ Jones \\& Forman 1984; Buote \\& Tsai 1996; Schuecker et al.\\ 2001) and the angular and line--of--sight velocity distribution of cluster galaxies (e.g.\\ Geller \\& Beers 1982; Dressler \\& Shectman 1988; West \\& Bothun 1990). These diagnostics have often been used as surrogates for a direct tracer of the underlying DM distribution. The major drawback of this approach is the requirement to assume a relationship between the luminous and dark matter distributions (e.g.\\ that the ICM is in hydrostatic equilibrium with the DM potential) -- it is precisely these assumptions that require detailed testing. This issue is further aggravated by the expectation that cluster mass distributions are DM dominated on all but the smallest scales (e.g.\\ Smith et al.\\ 2001; Sand, Treu \\& Ellis 2002; Sand et al.\\ 2004). Gravitational lensing offers a solution to much of this problem, in that the lensing signal is sensitive to the total mass distribution in the lens, regardless of its physical nature and state. Detailed study of gravitational lensing by massive clusters is therefore an important opportunity to gain an empirical understanding of the distribution of DM in clusters. Indeed, early comparisons between X--ray and lensing--based mass measurements revealed a factor 2--3 discrepancy between X--ray and strong--lensing--based cluster mass estimates (e.g.\\ Miralda--Escud\\'e \\& Babul 1995; Wu \\& Fang 1997), although the agreement between weak--lensing and X--ray measurements was generally better, albeit within large uncertainties (e.g.\\ Squires et al.\\ 1996, 1997; Smail et al.\\ 1997). The simplifying assumptions involved in the X--ray analysis were soon identified as the likely dominant source of this discrepancy; this was confirmed by several authors (e.g.\\ Allen 1998; Wu et al.\\ 1998; Wu 2000). In summary, X--ray and lensing mass measurements for the most relaxed clusters agree well if the multi--phase nature of the ICM in cool cores (e.g.\\ Allen et al.\\ 2001) is incorporated into the X--ray analysis. The situation is more complex in more dynamically disturbed clusters, with larger discrepancies being found at smaller projected radii. The origin of the X--ray versus lensing mass discrepancy in clusters that do not contain a cool core is generally attributed to the simplifying equilibrium and symmetry assumptions of the X--ray analysis. For that reason, most modern X--ray cluster analyses that involve measuring cluster mass using only X--ray data understandably concentrate on relaxed, cool core clusters (e.g.\\ Allen et al.\\ 2001). An important caveat to adopting lensing as the tool of choice to measure cluster mass is that lensing actually constrains the projected mass distribution along the line--of--sight to the cluster. The addition of three--dimensional information into lensing studies may therefore be important before final conclusions are drawn. For example Czoske et al.'s (2001; 2002) wide--field redshift survey of Cl\\,0024$+$1654 at $z{=}0.395$ revealed that this previously presumed relaxed strong--lensing cluster (e.g.\\ Smail et al.\\ 1996; Tyson et al.\\ 1998) cluster is not relaxed, and appears to have suffered a recent merger along the line--of--sight (see also Kneib et al.\\ 2003). Early gravitational lensing studies of galaxy clusters concentrated on individual clusters selected because of their prominent arcs (e.g.\\ Mellier et al.\\ 1993; Kneib et al.\\ 1994, 1995, 1996; Smail et al.\\ 1995a, 1996; Allen et al.\\ 1996; Tyson et al.\\ 1998). This ``prominent arc'' selection function was vital to development of the techniques required to interpret the gravitational lensing signal (e.g.\\ Kneib 1993). However it also made it difficult to draw conclusions about galaxy clusters as a population of astrophysical systems from these studies. Smail et al.\\ (1997) used the \\emph{Hubble Space Telescope (HST)} to study a larger sample of optically rich clusters, selected originally for the purpose of studying the cluster galaxies. As X--ray selected samples became available in the late--1990's, Luppino et al.\\ (1999) also searched for gravitational arcs in ground--based imaging of 38 X--ray luminous clusters. The broad conclusions to emerge from these pioneering studies were that to use gravitational lensing to learn about clusters as a population, a selection function that mimics mass--selection as closely as possible, and the high spatial resolution available from \\emph{HST} imaging are both key requirements. We are conducting an \\emph{HST} survey of an objectively selected sample of ten X--ray luminous (and thus massive) clusters at $z{\\simeq}0.2$ (Table~\\ref{sample}, \\S\\ref{design}). Previous papers in this series have presented (i) a detailed analysis of the density profile of A\\,383 (Smith et al.\\ 2001), (ii) a search for gravitationally--lensed Extremely Red Objects (EROs -- Smith et al.\\ 2002a) and (iii) near--infrared (NIR) spectroscopy of ERO\\,J003707, a multiply--imaged ERO at $z{=}1.6$ behind the foreground cluster A\\,68 (Smith et al.\\ 2002b). This paper describes the gravitational lensing analysis of all ten clusters observed with \\emph{HST} and uses the resulting models of the cluster cores to measure the mass and structure of the clusters on scales of $R{\\le}500\\,{\\rm kpc}$. We also exploit archival \\emph{Chandra} observations and NIR photometry of likely cluster galaxies to compare the distribution of total mass in the clusters with the gaseous and stellar components respectively. This combination of strong--lensing, X--ray and NIR diagnostics enable us to quantify the prevalence of dynamical immaturity in the X--ray luminous population at $z{\\simeq}0.2$ and to calibrate the high--mass end of the cluster mass--temperature relationship. We outline the organization of the paper. In \\S\\ref{design} we describe the survey design and sample selection. We then describe the reduction and analysis of the optical data in \\S\\ref{optical}, comprising the \\emph{HST} imaging data (\\S\\ref{hst}) and new spectroscopic redshift measurements for arcs in A\\,68 and A\\,2219 (\\S\\ref{spec}). The end--point of \\S\\ref{optical} is a definition of the strong-- and weak--lensing constraints available for all ten clusters. We apply these constraints in \\S\\ref{modelling}, to construct detailed gravitational lens models of the cluster potential wells; the details of the modelling techniques are described in the Appendix, and the process of fitting the constraints in each cluster are described in \\S\\ref{modelling}. We then complement these gravitational lensing results with observations of the X--ray emission from the clusters' ICM, drawn from the \\emph{Chandra} data archive (\\S\\ref{xray}). The main results of the paper are then presented in \\S\\ref{results}, including measurements of the mass and maturity of the clusters and a detailed study of the cluster scaling relations. We discuss the interpretation of the results in \\S\\ref{discussion} and briefly assess their impact on attempts to use clusters as cosmological probes. Finally, we summarize our conclusions in \\S\\ref{conclusions} We assume a spatially flat universe with $H_0{=}50{\\rm km\\,s^{-1}Mpc^{{-}1}}$ and $q_0{=}0.5$; in this cosmology $1''{\\equiv}4.2\\,{\\rm kpc}$ at $z{=}0.2$. Our main results are insensitive to this choice of cosmology, for example, the cluster mass measurements would be modified by ${\\ls}10\\%$ if we adopted the currently favored values of $\\Omega_{\\rm M}{=}0.3$, $\\Omega_\\Lambda{=}0.7$, $H_0{=}65{\\rm km\\,s^{-1}Mpc^{{-}1}}$. We also adopt the complex deformation, $\\vec\\tau{=}\\tau_x{+}i\\tau_y{=}|\\vec\\tau| e^{2i\\theta}$, as our measure of galaxy shape when dealing with the weak lensing aspects of our analysis, where $\\tau{=}(a^2{+}b^2)/2ab$ and $\\theta$ is the position angle of the major axis of the ellipse that describes each galaxy. We define the terms ``ellipticity'' to mean $\\tau$ and ``orientation'' to mean $\\theta$. All uncertainties are quoted at the 68\\% confidence level. ", "conclusions": "We have undertaken a comprehensive study of the distribution of mass in ten X--ray luminous ($L_X{\\ge}8{\\times}10^{44}{\\rm erg \\,s^{-1}}$[0.1--2.4\\,keV]) galaxy clusters at $z{=}0.21{\\pm}0.04$. The cornerstone of our analysis is a suite of detailed gravitational lens models that describe the distribution of total mass in the cluster cores. These models are constrained by the gravitational lensing signal detected in high--resolution \\emph{HST}/WFPC2 imaging of the clusters, including numerous multiply--imaged and weakly--sheared background galaxies. Analysis of archival \\emph{Chandra} observations complements the lensing analysis and enables us to relate the total mass and structure of the clusters to the thermodynamics of the intra--cluster medium. We re--cap the key results. \\setcounter{fred}{0} \\begin{list}{(\\roman{fred})}{\\usecounter{fred}\\setlength{\\itemindent}{0mm}\\setlength{\\labelwidth}{15mm}\\setlength{\\labelsep}{2mm}\\setlength{\\leftmargin}{7mm}} \\setlength{\\itemsep}{1.0mm} \\item Five of the ten clusters contain spectroscopically confirmed strong gravitational lensing, i.e.\\ multiply--imaged background galaxies. These five clusters comprise: A\\,68 (Smith et al.\\ 2002b; \\S\\ref{spec}), A\\,383 (Smith et al.\\ 2001; Sand et al.\\ 2004), A\\,963 (Ellis et al.\\ 1991), A\\,2218 (Pell\\'o et al.\\ 1992; Ebbels et al.\\ 1998; Ellis et al.\\ 2002) and A\\,2219 (\\S\\ref{spec}). \\item Of the remaining five clusters, two contain relatively unambiguous examples of strong lensing for which spectroscopic redshifts are not yet available (A\\,267 and A\\,1835). The other three clusters, A\\,209, A\\,773 and A\\,1763 do not contain any obviously multiply--imaged galaxies, however the optical richness and massive nature of A\\,773 imply that this cluster may well contain strong lensing that has yet to be uncovered. \\item Based on our search for strong--lensing in these clusters down to a surface brightness limit of $\\mu_{702}{\\simeq}25$\\,mag\\,arcsec$^{{-}2}$, we therefore put a firm lower limit on the fraction of the cluster sample that have a central projected mass density in excess of the critical density required for gravitational lensing of $50\\%$. Including A\\,267 and A\\,1835 increases this limit to $70\\%$. \\item We use the strong-- and weak--lensing signals to constrain parametrized models of the cluster potential wells, and from these models compute maps of the total projected mass in the cluster cores. Spatial analysis of these maps reveals that four of the clusters form a homogeneous sub--sample with very high central mass fractions (${\\Mcen}/{\\Mtot}{>}0.95$). The remaining six are strongly heterogeneous, with central mass fractions in the range $0.4{\\le}{\\Mcen}/{\\Mtot}{\\le}0.9$. The central mass fraction of ${\\Mcen}/{\\Mtot}{\\simeq}0.95$ that divides these two populations corresponds to a $K$--band central luminosity fraction of $L_{\\rm K,BCG}/L_{\\rm K,tot}{\\sim}0.5$. \\item All of the six low central mass fraction clusters have an irregular, but not obviously bi/tri--modal X--ray morphology. Four of the six are constrained by the current lensing data to have a bi/tri--modal mass morphology (A\\,68, A\\,773, A\\,2218, A\\,2219). The other two (A\\,209 and A\\,1763) may be merging in the plane of the sky and thus any multi--modality in their mass distributions is not well--sampled by our WFPC2 pencil--beam survey of the cluster cores. \\item Three of the four high central mass fraction clusters also have relaxed X--ray morphologies. The remaining cluster (A\\,267) has a disturbed X--ray morphology, with a ${\\sim}90$\\,kpc offset between its centers of X--ray emission and mass. The distribution of mass in this cluster may therefore be more complex than the single dark matter halo (plus cluster galaxies) that the current data are able to constrain. \\item Combining all of the information available to us, we define ``relaxed'' clusters to be those which appear relaxed in all available diagnostics, with the exception that we do not require a cool core. Quantitatively relaxed clusters therefore have a single cluster--scale DM halo in their lens model ($N_{\\rm DM}{=}1$), a high central mass fraction (${\\Mcen}/{\\Mtot}{\\ge}0.95$) and central $K$--band luminosity fraction ($L_{K,{\\rm BCG}}/L_{K,{\\rm tot}}{\\gs}0.5$), no evidence for an offset between the X--ray emission and the center of mass ${\\drpeak}{<}4{\\rm kpc}$) and the X--ray morphology is either circular, or mildly elliptical. The unrelaxed clusters do not meet at least one of these criteria. \\item Applying these criteria to the cluster sample, we conclude that seven of the ten clusters are dynamically immature, i.e.\\ unrelaxed (A\\,68, A\\,209, A\\,267, A\\,773, A\\,1763, A\\,2218, A\\,2219) and three are relaxed (A\\,383, A\\,963, A\\,1835); thus, formally $70{\\pm}20\\%$ of X--ray luminous cluster cores at $z{=}0.2$ are unrelaxed. \\item We detect a factor of three more scatter in the mass--temperature plane than predicted by Evrard, Metzler \\& Navarro (1996), implying that great care should be exercised when using such relations to convert cluster temperature functions to mass functions in pursuit of cosmological parameters. We also consider a number of key uncertainties that may artificially inflate our estimate of the scatter. This exercise suggests that approximately one third of the scatter detected in this study may be due to issues related to the small field--of--view of our WFPC2 observations. \\item The scatter in the mass--temperature plane is asymmetric, presenting evidence of structural segregation. The normalization of the mass--temperature relation for unrelaxed (dynamically immature) clusters is $40\\%$ hotter than for relaxed clusters at 2.5--${\\sigma}$ significance. This result is consistent with recent simulations of cluster--cluster mergers (Ricker \\& Sarazin 2001; Randall, Sarazin \\& Ricker 2002), implying that merger induced temperature boosts may be the dominant factor behind the hotter normalization of unrelaxed systems. \\end{list} In summary, this study is the first of its kind, exploiting detailed strong--lensing constraints on the distribution of mass in X--ray luminous cluster cores, complemented by X--ray spectro--imaging with \\emph{Chandra}. The high frequency of dynamical immaturity, coupled with the structural segregation of the clusters in the mass--temperature plane have profound implications for our understanding of how clusters form and evolve. Perhaps of greatest importance at this time is the implications of these results for using clusters to constrain the cosmological parameters, ${\\Omega}_M$, ${\\sigma}_8$ and the dark energy equation--of--state parameter $w$. In a companion paper we demonstrate that inadequate knowledge of the cluster selection function can lead to $20\\%$ systematic errors in cluster--based measurements of ${\\sigma}_8$. Turning to $w$, forthcoming Sunyaev--Zeldovich Effect experiments designed to detect and measure the mass of clusters out to high redshifts, using mass--temperature scaling relations may be compromised by unidentified and/or poorly calibrated astrophysical systematic uncertainties (see also Majumdar \\& Mohr 2003). Our future program will build on these results in three ways. First, we aim to overcome the principal uncertainties in the current work: small number statistics and tiny field--of--view. Wide--field space--based imaging of a statistically complete sample of clusters would be essential to achieve this goal. Second, detailed comparison of selection effects and measurement techniques between theoretical and observational studies will enable more detailed and rigorous comparison between observational and synthetic datasets. Finally, we are gathering \\emph{HST}/ACS imaging of an identically selected sample of clusters at $z{\\simeq}0.55$ drawn from the MACS sample (Ebeling et al.\\ 2001b). We will combine these data with observations in the X--ray pass--band and compare the results to those found here. The ${\\sim}3$\\,Gyr difference in look--back--time between $z{=}0.2$ and $z{=}0.55$ will enable us to search for evolutionary trends in the most massive clusters." }, "0403/astro-ph0403294_arXiv.txt": { "abstract": "Using the spectroscopic sample of the SDSS DR1 we measure how gas was transformed into stars as a function of time and stellar mass: the baryonic conversion tree (BCT). There is a clear correlation between early star formation activity and present-day stellar mass: the more massive galaxies have formed about 80\\% of their stars at $z>1$, while for the less massive ones the value is only about 20\\%. By comparing the BCT to the dark matter merger tree, we find indications that star formation efficiency at $z>1$ had to be about a factor of two higher than today ($\\sim 10\\%$) in galaxies with present-day stellar mass larger than $2 \\times 10^{11}M_\\odot$, if this early star formation occurred in the main progenitor. Therefore, the LCDM paradigm can accommodate a large number of red objects. On the other hand, in galaxies with present-day stellar mass less than $10^{11}$ M$_{\\odot}$, efficient star formation seems to have been triggered at $z \\sim 0.2$. We show that there is a characteristic mass (M$_* \\sim 10^{10}$ M$_{\\odot}$) for feedback efficiency (or lack of star formation). For galaxies with masses lower than this, feedback (or star formation suppression) is very efficient while for higher masses it is not. The BCT, determined here for the first time, should be an important observable with which to confront theoretical models of galaxy formation. ", "introduction": "In the current paradigm for galaxy formation, galaxies form in cold dark matter halos, which evolve from small, primordial, Gaussian fluctuations, by gravitational instability. This mechanism fits well in the successful LCDM picture which correctly describes the Universe from the Cosmic Microwave Background at $z=1088$ to local galaxy clustering \\citep{SpergelWMAP03,Percival2dF03}. One of the strong predictions of this galaxy formation paradigm is the typical redshift of dark matter halo formation (i.e. virialization) as a function of halo mass and cosmological parameters (e.g., \\citet{ST04}). Given that cosmological parameters have been tightly constrained (e.g., \\cite{SpergelWMAP03}), we can reconstruct the average dark halo formation history as a function of mass (e.g., \\citet{PS74,ST99,ST04}). Naively, the dark matter halo collapse, i.e. virialization, should trigger baryonic gas transformation into stars; in addition, subsequent dark matter mergers should produce star formation episodes. We will show that this simple model is not in agreement with observations. All we can observe is the integrated light of galaxies' stellar population; thus, to compare the theory prediction for the dark matter halos with observations, the process of how baryons are transformed into stars needs to be simulated, either through semi-analytical recipes or by means of hydro-dynamic N-body simulations. Since no theory of star formation has been yet established, we do not have a fundamental theory that allows us to compute from basic principles the star formation efficiency. Further, complications arise from other phenomena, such as feedback by stars and AGN that prevent the formation of giant molecular clouds and therefore reduce star formation efficiency. Given the complexity of the task of learning about galaxy formation from numerical simulations, here we take the complementary approach of placing new observational constraints on the stellar assembly history as a function of galaxy mass. In this paper we use about $10^5$ galaxies from the Sloan Digital Sky Survey Data Release 1 (SDSS DR1) to determine, for the first time, the amount of baryons that have been transformed into stars as a function of total stellar mass and time. This allows us to build the baryonic conversion tree (BCT), which can then be compared with the dark matter merger tree. We present such comparison and show how it can be used to compute the star formation efficiency and the relative importance of feedback (or lack of star formation). Our main findings are: \\begin{enumerate} \\item There is a clear correlation between total stellar mass of the galaxy and the fraction of gas transformed into stars at $z \\geq 1$. The larger the mass the larger the fraction of gas transformed into stars at high redshift. \\item If large galaxies need to be formed by $z \\sim 1$ in a ``monolithic'' fashion, as observations suggest (e.g., \\citet{Bower+92,PJDWSSDW98,Lilly+98,BE00,Im+02,Renzini03,Gao+03,Glazebrook+04}), high-redshift star formation efficiency needs to be much higher (about 10\\%) in massive galaxies than in less massive ones. This high star formation activity at early times means the build-up of stellar mass does not follow the hierarchical build-up of total mass. Stars could form in smaller objects (not in the main progenitor only) with lower efficiency, provided that the galaxy formation process has some way to synchronise star formation in disparate pieces. The reason being that the above mentioned galaxies at $z\\sim 1-2$ have stellar populations with almost no spread in their stellar ages and derived star formation histories consistent with very short times ($< 0.1$ Gyr, see Macarthy et al 2004). \\item There is an indication that major mergers are not the principal drivers of star formation. \\item We propose that a threshold for star formation for galaxies with masses $\\sim 10^{10}$ M$_{\\odot}$ can explain the findings above. The existence of such a threshold at low redshift is well documented in the literature \\citep{MK01}, and can be linked to feedback efficiency. Feedback, i.e. suppression of star formation activity, is expected to be very inefficient in massive galaxies, while efficient in less massive galaxies. \\end{enumerate} ", "conclusions": "We have determined for the first time the baryonic conversion tree for galaxies. We were able to do this with observations at $z<0.3$, using 96545 galaxies from the SDSS DR1 spectroscopic sample. In the hierarchical structure formation model, massive {\\em dark halos} form (i.e. virialize) later than smaller halos, from mergers of smaller units (e.g. \\citet{lc93,lc94,ljl03}). This model has been thoroughly tested against numerical cold dark matter simulations. Naively one could expect that the dark matter halo collapse should trigger baryonic gas transformation into stars; in addition subsequent dark matter mergers should produce star formation episodes. Instead, for the the stellar assembly history we find that: the more massive galaxies have old stellar population and massive, old elliptical galaxies are already in place at $z \\sim 1$. This has been known for a long time and sometimes it is referred to as ``down-sizing'', \\citep{CSC96}. We find that massive galaxies have transformed more gas into stars at higher redshift (in agreement with high z observations e.g., \\cite{kodama04}), and then star formation was suppressed, while less massive galaxies transform more gas into stars at low redshift. So one should not be surprised to see abundant red objects at high redshift in the LCDM paradigm, these objects can form in virialised halos if star formation efficiency is high. Indeed, \\cite{JFDTPN99} have shown that single-halo hydro-dynamical models would require an increased star formation efficiency for more massive galaxies, higher than the few percent found today in giant molecular clouds, in agreement with the value determined from the ``fossil record'' in the present work. Our findings, based only on observations at $z < 0.35$ (the ``fossil record''), are in agreement with a suite of independent, $z>1$, observations: observations of old elliptical galaxies at high redshift \\citep{D+96,Spinrad+97,NDJH03}, indications that elliptical galaxies are already formed at $z>1$ (e.g. \\citet{Bower+92,PJDWSSDW98,Lilly+98,BE00,Im+02,Renzini03,Gao+03,Glazebrook+04}). It also nicely explains the tightness of the colour-magnitude relation \\citep{Bower+92}. On the other hand, we find that small galaxies seem to accrete mass passively at early times and see a lot of their gas reservoir transformed into stars at late times. Since the probability distribution for dark halo mergers peaks at low redshift for massive halos and high redshift for small halos, we conclude that dark matter mergers and star formation are not correlated. We speculate that one possibility to explain the apparent and illusory anti-hierarchical nature of the stellar assembly history is the existence of a threshold for star formation: once the threshold is crossed all available baryons are turned into stars (``infall model'') and afterward galaxies approximatively evolve passively. The threshold is met at very early time for massive galaxies and a later time for less massive ones (see e.g. \\citet{HJ99}). A star formation threshold has been observed in disk galaxies by \\citet{MK01} and there has been some recent additional evidence from the formation of dust lanes in disk galaxies \\citep{DYB04} that this threshold may take place at $V_c \\sim 100$ km s$^{-1}$, in agreement with the findings of \\citet{VOJ02} and \\citet{kannappan04} who also found a transition at about $100$ km s$^{-1}$ for star formation efficiency. This $V_c$ value corresponds to the characteristic mass found here ($M_* \\sim 10^{10}$ M$_{\\odot}$), that defines the border between efficient and inefficient star formation. This characteristic mass has been related to feedback efficiency threshold (e.g., Dekel \\& Silk 1986, Dekel \\& Woo 2003 and references therein). As we do not yet have a fundamental theory for galaxy formation and given the complexity involved in studying the process with hydrodynamic-N body simulations, we hope that this new determination of the baryonic conversion history will be a useful observable to gauge galaxy formation models against.\\\\" }, "0403/astro-ph0403407_arXiv.txt": { "abstract": "We show how the viscous evolution of Keplerian accretion discs can be understood in terms of simple kinetic theory. Although standard physics texts give a simple derivation of momentum transfer in a linear shear flow using kinetic theory, many authors, as detailed by Hayashi \\& Matsuda 2001, have had difficulties applying the same considerations to a circular shear flow. We show here how this may be done, and note that the essential ingredients are to take proper account of, first, isotropy locally in the frame of the fluid and, second, the geometry of the mean flow. ", "introduction": "Accretion discs play a central role in a wide range of astronomical environments, mediating the gas flows in the vicinity of object as diverse as AGN, binary stars and protostars (Pringle, 1981). In an accretion disc, the predominant flow is a circular shear flow, with angular velocity $\\Omega(R)$ a function of radius $R$ from the central object. Accretion takes place because of the action of some form of dissipation which releases the free energy of the shear flow as heat, and so allows the disc material to fall deeper into the potential well of the central object. Simple physical energy arguments ({\\it e.g.} Lynden-Bell \\& Pringle, 1974) indicate that the dissipative process must take the form of a stress which transports angular momentum outwards. Because the free energy of a circular shear flow is zero only if $d\\Omega/dR = 0$, it follows that the relevant element of the stress tensor must be of the form \\begin{equation} T_{\\phi R} \\propto - d\\Omega/dR, \\end{equation} where $\\phi$ is the azimuthal coordinate. This can be deduced from the standard derivation of Navier-Stokes stress to be found in the fluid dynamics textbooks. However, in a recent paper, Hayashi \\& Matsuda (2001) have drawn attention to the fact that attempts to provide a physical explanation of the above result in terms of simple kinetic theory have resulted in failure. The simple kinetic explanation given in basic physics text books for the effect of viscosity on a simple linear shear flow, of the form ${\\bf u} = (0,U(x))$ in Cartesian coordinates, relies on the fact that the kinetic particles conserve linear momentum between collisions. Thus particles crossing some fiducial plane $x = x_o$ tend to mix up and smooth out the momentum distribution of the fluid and so give rise to a stress of the form $T_{yx} \\propto -dU/dx$. However, as Hayashi \\& Matsuda (2001) point out, it is in the application of these simple concepts to a circular shear flow that the problems seem to arise. The simple generalisation that, in a circular shear flow, the kinetic particles conserve angular momentum $j = R^2 \\Omega$ between collisions would imply, taken at face value, that the the movement of particles across some fiducial circle $R=R_o$ would tend to try to mix up and smooth out the distribution of angular momentum of the fluid, and thus that the stress would be proportional to $-dj/dR$ (see e.g. Madej \\& Paczynski 1977). From the arguments given above, this is clearly wrong. Not only would this predict a stress in the case when the shear $d\\Omega/dR$ is zero, but for a standard Keplerian accretion disc for which $j \\propto R^{1/2}$ it would transport angular momentum inwards rather than outwards. As detailed by Hayashi \\& Matsuda (2001) attempts to get round this and to produce the 'correct' result have only succeeded by making mathematical errors in the derivation. From all these problems, Hayashi \\& Matsuda (2001) conclude that although what they call the derivation 'with mathematical rigour' ({\\it i.e.} the usual Navier-Stokes argument) gives the correct answer, in order to obtain the correct answer using kinetic theory one must take account of such complications as Coriolis force. In this paper, we shall show that, although it is obviously necessary to include Coriolis force if one works in a frame co-rotating with the flow, one can obtain the correct result from straightforward kinetic theory in the inertial frame. Before we do so, it is instructive to return to the 'mathematical' relationship between stress and strain derived in the standard fluid textbooks (see, for example, Batchelor, 1967, Section 3.3). The basic point we wish to make is that the standard Navier-Stokes expression for momentum transfer ({\\it i.e.} stress) is based on a simple {\\it physical} argument. The argument may involve the use of tensors, which physicists tend to meet in courses on mathematical methods, but the argument itself is not a mathematical one. The point is that stress (momentum transfer) in a fluid (or in a solid) can be expressed as a second order tensor. This is a physical (and not a mathematical) statement, in exactly the same sense that the statement that a velocity is a first order tensor (i.e. a vector) is a physical, and not a mathematical, statement.\\footnote{The mathematical complications come in when one has to specify a particular tensor (or vector) by specifying the coordinate representation of that tensor in a particular coordinate frame, and get worse when one has to specify the representation of that same tensor in some other coordinate frame.} For a fluid, the physical {\\it Ansatz} is simply that the stress tensor must be physically related to the rate of strain tensor (which is a second order tensor which derives from the first derivatives of the velocity field, and thus incorporates information about the shear). The simplest assumption is that relationship between these two tensors is through a (fourth order) tensor which is isotropic. It is this assumption of the isotropy of the relationship, which is based on the physical assumption that the fluid itself is isotropic, which gives rise to the standard Navier-Stokes expression for the viscosity.\\footnote{This simplest assumption is not necessarily the correct one. For example it is possible that the viscosity might depend on the absolute orientation of the shear in some inertial frame ({\\it e.g.} the stress due to convection in a rotating medium -- see Kumar, Narayan \\& Loeb 1995, and references therein; the stress due to non-isotropic mixing -- Bretherton \\& Turner, 1968; or the stress induced by the magneto-rotational instability -- Torkelsson {\\it et. al.} 2000).} It is important to realise that these tensors ({\\it i.e.} scalars, vectors, second order tensors) exist as physical quantities, independent of any coordinate system. The statement that there is a relationship between two of them is a physical statement. It is only when one has to calculate a particular element of the stress tensor, for example corresponding to linear momentum transfer in a linear shear flow, or to the angular momentum transfer in a circular shear flow, that one has to evaluate coordinate dependent expressions, which can get mathematically complicated. But when one does this, one finds that the flux of linear momentum in a linear shear flow just depends on $-dU/dx$, and that the angular momentum flux in a circular shear flow just depends on $-d\\Omega/dR$. However, this result also enables us to draw another conclusion. The reason for the difference between the terms $dU/dx$ and $d\\Omega/dR$ is due solely to the difference between the coordinate systems. That is, it comes from geometry alone. This implies that when looking for differences in derivations for simple kinetic theory formulae between the linear and circular shear flows, we need only concern ourselves with geometry, and not with dynamical complications such as Coriolis force. In addition, we also need to take note of the fact that the Navier-Stokes expression does depend critically on assumptions about isotropy of the fluid. Thus we should expect to have to make a similar assumption about the properties of our kinetic particles. ", "conclusions": "We have shown that it is possible to obtain the correct expression for the momentum transfer in a both a linear and a circular shear flow using simple kinetic theory. This correct expression implies that an isotropic viscosity transfers momentum {\\it down} an angular velocity gradient. We have noted that it is essential to take proper account of both the geometry and the fluid's isotropy. \\begin{figure} \\vspace{2pt} \\epsfig{file=fig3.eps,width=8.5cm,height=7.cm} \\caption{Velocity components for two particles, $1$ and $2$, in a linear shear flow originating at emission points located at $ \\pm \\alpha$. Key as in Figure 1.} \\end{figure} Our analysis allows us to understand the flaw in the simple heuristic argument applied to a circular shear flow (described in Section 1), which leads to the conclusion that since angular momentum is conserved along particle trajectories, it is the gradient in angular momentum which particle mixing tends to smooth out. Particles arrive at the reference patch, at point S, with momentum along the streamline ($y-$momentum) that derives from two sources -- (i) the $y-$component of the random emission velocity (assumed isotropic in the local rest frame of the emission point) and (ii) the $y-$component of the streaming velocity of the emission point relative to the reference patch. \\begin{figure} \\vspace{2pt} \\epsfig{file=fig4.eps,width=8.5cm,height=7.cm} \\caption{Velocity components for two particles, $1$ and $2$, in a circular shear flow originating at emission points located at $ \\pm \\alpha$. Key as in Figure 1.In this example, $R \\Omega^\\prime + \\Omega =0 $ (i.e. the speed of the flow is everywhere constant) but nevertheless the total velocity vector (open arrowhead) is larger for particle $1$, implying that there is a net flux of $y$ momentum at S.} \\end{figure} The usual argument given to explain the momentum transfer in a {\\it linear} shear flow is simply that (ii) has the same sign for all the emission points on one side of the reference patch and so that momentum is added to, or subtracted from, the reference patch just depending on the sign of the velocity gradient in the mean flow. In fact, there is another contribution, equal in magnitude and sign to that described above, which may be understood by considering the particles that are incident at the reference patch from $\\pm \\alpha$. From Figure 3 it may be seen that the $x-$velocity (and hence flux) of particles on the $+ \\alpha$ side exceeds that on the $- \\alpha$ side ({\\it i.e.} there is a greater flux of particles at the reference patch whose emission is prograde than retrograde). Since the $y-$momentum from (i) far exceeds that from (ii), this slight asymmetry in the fluxes of prograde and retrograde particles produces a net momentum flux that is equal to the more obvious source of momentum flux described above. We may now apply the same considerations to the {\\it circular} shear flow, where once again there are the two contributions to the $y-$momentum -- (i) and (ii) described above. The important difference now is that the mean streaming velocity is no longer in the $y-$direction. The geometry of the circular arc ensures that this contributes a positive (negative) $x-$velocity for $\\alpha > 0 \\: (\\alpha < 0)$, respectively. As may be seen from Figure 4, particles that are incident from $+ \\alpha$ (prograde particles) have a larger amplitude of both $v_x$ and $v_y$ than those from $- \\alpha$. This ensures that the mean $y-$velocity of particles arriving at the reference patch from a particular emission streamline is not equal to the mean $y-$velocity of particles on that emission streamline. This disparity is because of the relative boost in the arrival flux of particles whose random velocity is prograde in the frame of the emission streamline, compared with those that are retrograde. This now allows us to understand the behaviour of a Keplerian accretion disc at a qualitative level. The mean angular momentum of particles orbiting at radii less than that of the reference patch is less than that of the reference patch. If the distribution of angular momenta of the particles arriving at the reference patch merely reflected the distribution of angular momenta of particles on their parent emission streamlines, this would imply that the arrival of particles at the reference patch from smaller radii should exert a spindown torque. Instead, the relative boost in the {\\it arrival rate} of particles that are emitted in the prograde direction, ensures that the average angular momentum of the particles arriving at the reference patch exceeds the average at the parent streamline. In the Keplerian case, this relative boost in the arrival rate of the prograde particles is enough to reverse the sign of angular momentum transfer. Thus, particles arriving at the reference patch from smaller radii exert a spin up torque, as required." }, "0403/astro-ph0403631_arXiv.txt": { "abstract": "We have used an antenna temperature thresholding algorithm on the Bell Laboratories $\\thco$ Milky Way Survey to create a catalog of 1,400 molecular clouds. Of these, 281 clouds were selected for having well-determined kinematic distances. The scaleheight, luminosity, internal velocity dispersion, and size of the cloud sample are analyzed to show that clouds smaller than $\\sim 10^{5.5} \\, \\msol$ have a scaleheight which is about 35 pc, roughly independent of cloud mass, while larger clouds, the Giant Molecular Clouds, have a reduced scaleheight which declines with increasing cloud mass. ", "introduction": "Since star formation occurs in molecular clouds, the formation of molecular clouds is an essential first step to the star formation process. The overall evolution of galactic metallicity implies that hot ejecta from old stars must somehow return to the cold molecular phase in order to form new stars, and that the timescale for this process is short compared to the age of the Galaxy \\citep[e.g.][]{freeman02}. Giant Molecular Clouds (GMCs) are the largest concentrations of molecular material, amounting to $2 \\times 10^5 \\msol$ or more in a region about 50 pc in size \\citep{stark78}. GMCs are strongly concentrated to spiral arms \\citep{stark79c,lee01}, suggesting that they are transient phenomena: hundreds of thousands of solar masses of molecular material collect during the passage of a spiral density wave, and then dissipate some thirty million years later. Thousands of stars are formed during the lifetime of the GMC, including the generators of the giant \\ion{H}{2} regions that trace spiral structure. How molecular clouds form, and why they form, has been the subject of ongoing theoretical investigation \\citep{elmegreen00,pringle01}. Computer simulations of cloud formation have become increasingly complex and realistic \\citep{hartmann01,zhang02}. The cloud formation process gathers and concentrates interstellar matter. However this happens, we expect that the random velocity of the resulting cloud would be less than the velocity dispersion of the precursor material, so that the velocity dispersion of GMCs as a class would be less than that of smaller interstellar clouds. This hypothesis is amenable to observational test. The purpose of this {\\em Letter} is to describe such an observational test, and then to quantify the effect in a way which may make for useful comparisons with theory. In \\S 2, we take the Bell Laboratories $\\thco$ Survey data and use a brightness temperature thresholding algorithm to create a catalog of molecular clouds. We then use our knowledge of Milky Way structure in order to select a subset of clouds with well-determined kinematic distances. This allows us to determine the physical size of the cloud and its distance above the galactic plane. These data are analyzed in \\S 3 to show that the scaleheight of large clouds is less than that of smaller clouds. ", "conclusions": "The data show that small molecular clouds have a scaleheight which is approximately independent of cloud size. Larger clouds, the GMCs, have a scaleheight which falls off with mass. This has been demonstrated with a relatively small sample of clouds, but it is a clean sample, chosen from a large-scale survey by an algorithmic method. The implication is that the Giant Molecular Clouds, those clouds which are concentrated in the spiral arms of the Galaxy, are also concentrated to the galactic plane. We can understand this as a manifestation of the molecular cloud formation process. Atomic gas clouds, with a scaleheight of $\\sim 100 \\, \\mathrm{pc}$ and a velocity dispersion $\\sim 12 \\, \\kms$, condense out of the diffuse atomic gas. Their cores become molecular and increasingly more condensed, resulting in a population of high-latitude ($\\sim 70 \\, \\mathrm{pc}$), low mass ($M \\ls 100 \\, \\msol$) partially-molecular clouds with a dispersion $\\sim 8 \\, \\kms$ \\citep{malhotra94b,dame94}. These clouds become larger, more centrally condensed, and more bound, with only a slight reduction in scaleheight and velocity dispersion, to become molecular clouds ($ 100 \\, \\msol \\ls M \\ls 10^5 \\, \\msol$), with dispersion $\\sim 7 \\, \\kms$ \\citep{stark89b}. These clouds can form stars, but are uniformly distributed throughout the galactic disk, and may survive for many galactic rotations. The largest clouds, the GMCs, are rapidly assembled by the passage of a spiral arm. This process is dissipative, different from the slow addition of material that forms smaller molecular clouds. It results in a significant loss of random velocity per unit mass, and the resulting GMCs are found at the galactic midplane, in the spiral arm. The ensuing formation of massive stars destroys the cloud, fragmenting it into stars, ionized gas, and small clouds before the next interarm passage." }, "0403/astro-ph0403361_arXiv.txt": { "abstract": "Evolution of first population of massive metal-free binary stars is followed. Due to the low metallicity, the stars are allowed to form with large initial masses and to evolve without significant mass loss. Evolution at zero metallicity, therefore, may lead to the formation of massive remnants. In particular, black holes of intermediate-mass ($\\sim 100-500$\\ M$_\\odot$) are expected to have formed in early Universe, in contrast to the much lower mass stellar black holes ($\\sim 10$\\ M$_\\odot$) being formed at present. Following a natural assumption, that some of these Population III stars have formed in binaries, the physical properties of first stellar binary black holes are presented. We find that a significant fraction of such binary black holes coalesces within the Hubble time. We point out that burst of gravitational waves from the final coalescences and the following ringdown of these binary black hole mergers can be observed in the interferometric detectors. We estimate that advanced LIGO detection rate of such mergers is at least several events per year with high signal to noise ratio ($\\gtrsim 10$). ", "introduction": "The properties of Population III stars have stirred a lot of interest in the recent years. It has been realized that zero metallicity stars with masses up to several hundred of solar masses are stable \\citep{2001ApJ...550..890B}. \\citet{2002ApJ...567..532H} estimated black hole (BH) masses formed in the evolution of high-mass metal-free stars. For initial masses above $40$\\ M$_\\odot$ the remnant is a BH with the mass essentially the same as the progenitor with the exception of stars with initial masses between $140$ and $260$\\ M$_\\odot$ which undergo a pair instability supernova (SN) explosions and leave no remnant at all. \\citet{2003ApJ...591..288H} found that these conclusions hold also for low, non-zero metallicity stars. Numerical studies of collapse and fragmentation of a metal free gas in the early Universe \\citep{1999ApJ...527L...5B,2002ApJ...564...23B,2003ApJ...589..677O} indicate that the initial mass function (IMF) of Population III stars is top heavy, and might be bimodal with the high mass peak around $100$\\ M$_\\odot$ \\citep{1998MNRAS.301..569L,2001ApJ...548...19N,2003PASP..115..763C}. Every known stellar system contains a considerable fraction of binaries. Therefore it seems quite natural to allow for a possibility that some fraction of massive Population III stars formed in binaries as well, and then investigate their subsequent evolution. There is a possibility that these systems form massive black hole black hole (BH-BH) binaries which in turn may be observable in gravitational waves by the interferometric detectors. In this paper we consider the properties of Population III binaries with the initial component masses in the range $100-500$\\ M$_\\odot$. In \\S\\,2 we describe the model of evolution of metal-free systems leading to formation of massive BH-BH binaries. In \\S\\,3 we present characteristic properties of the BH-BH population and discuss the observability of their mergers. Conclusions and summary of results are given in \\S\\,4. ", "conclusions": "The principal result of the calculations is presented in Figure~\\ref{rsn}. The expected coalescence rate of Population III intermediate-mass BH-BH binaries is high. We predict that advanced LIGO should observe above a thousand strong ($S/N \\gtrsim 10$) events per year. What are the uncertainties of this prediction? In our calculation we have assumed that Population III stars were formed at redshifts $z$ from $30$ to $10$ out of $f_{\\rm mass} = 10^{-3}$ of the baryon mass contained in the Universe and that the binary fraction of the initial population was $f_b=0.1$. The exact length and duration of the Population III star formation era does not affect the rate calculation. However, the rate scales linearly with $f_{\\rm mass}$ and $f_b$, and although we have chosen rather conservative values for these two quantities, the rate may decrease if the adopted values were significantly lowered. The IMF of Population III stars is another unknown. Numerical investigations show that the IMF leans towards massive stars. We assumed rather steep IMF (with the slope $\\alpha=-2$), to assure that we do not overproduce highest mass, and therefore the easiest to detect, BH-BH binaries. As it turns out, the change of IMF slope does not significantly alter the detection rate. Farther steepening of the IMF ($\\alpha=-3$) decreases the expected rate by half, while flatter IMF assumption ($\\alpha=-1$) increases the expected rate by a factor of two. Note that we maintain a flat IMF between $1\\,{\\rm M}_\\odot$ and $100\\,{\\rm M}_\\odot$ when calculating $f_{sim}$ -- the fraction of stars that we simulate out of the total population. The observed rate scales linearly with $f_{sim}$. We also fixed the maximum mass of the Population III stars to $500\\,{\\rm M}_\\odot$. Had we allowed for possibility of star formation with the higher mass the rate would increase. The estimate of the lifetime of the black hole binaries can be strongly affected if the binaries are hardened by interactions in dense stellar environments. Our simulations show that a large number of black hole binaries should be formed with the lifetimes in excess of the Hubble time. Interactions in dense systems may significantly shorten their lifetimes. We demonstrated above that hardening does not affect strongly the expected rate. Another possibility arises that the interactions disrupt some of BH-BH binaries. This would deplete the population of wide systems. None of the two effects, unless operating on extremely short timescales, can affect much systems with shortest coalescence timescales. Therefore, a combination of the two effects may constrain the mergers of Population III BH-BH binaries to large redshifts. In order to estimate the $S/N$ for advanced LIGO we have used the formulae of \\citet{1998PhRvD..57.4535F}. The predicted values of $S/N$ may still change when more realistic noise curves and binary gravitational wave form templates are known. Because of large masses of the binaries considered in this work the changes in the low-frequency range are most important. The typical ringdown frequency is $\\nu_{qnr} \\approx 90 (300{\\rm M}_\\odot /M) $\\,Hz, so the rate is most sensitive to the detector performance in the low frequency region. The scaling of the rate with the change of $S/N$ normalization can be read off Figure~\\ref{rsn}. The influence of the evolutionary model assumptions on detection rate was tested. We changed the CE efficiency (increase/decrease by factor of 2); altered evolution through stable RLOF phases from fully conservative to non-conservative cases; changed the specific angular momentum of the matter leaving the system during RLOF (increase/decrease by factor of 2) and varied the critical mass ratio over which the dynamical instability develops (from $q_{\\rm crit}=2$ in standard model to $q_{\\rm crit}=1-3$). The detection rate was decreased at most by a factor of 3 in the above models. Therefore, the predicted rate does not depend strongly on the binary evolution within the model assumptions. Summarizing, in our calculations we have tried to use rather conservative assumptions in order not to overestimate the detection rate of Population III BH-BH mergers. If several of the assumptions and values of the model parameters are changed within reasonable limits, we still obtain a significant detection rate. Even if the rate predicted for our already conservative model is decreased by 2-3 orders of magnitude to allow for different aforementioned uncertainties, we still are left with several strong events per year for advanced LIGO detector. For initial LIGO phase the detection rate falls below one event per year. We have shown that Population III stars may lead to formation of a large number of binaries containing intermediate-mass black holes. A significant fraction of such systems has coalescence times smaller than the Hubble time. Coalescences of intermediate-mass BH-BH binaries should be detectable by advanced interferometric gravitational wave detectors during the ringdown phase, and possibly also merger phase, provided that accurate templates are available. Given the large expected rate of observed coalescences such events could be the primary targets for LIGO burst \\vspace*{-0.35cm} search." }, "0403/astro-ph0403682_arXiv.txt": { "abstract": "Using the Far Ultraviolet Spectroscopic Explorer (FUSE) we have obtained 87 spectra of 57 low-redshift ($z<0.15$) active galactic nuclei (AGN). This sample comprises 53 Type 1 AGN and 4 Type 2. All the Type 1 objects show broad \\ovi\\ $\\lambda 1034$ emission; two of the Type 2s show narrow \\ovi\\ emission. In addition to \\ovi, we also identify emission lines due to \\ciii\\ $\\lambda 977$, \\ciii\\ $\\lambda 991$, \\siv\\ $\\lambda\\lambda 1062,1072$, and \\heii\\ $\\lambda 1085$ in many of the Type-1 AGN. Of the Type 1 objects, 30 show intrinsic absorption by the \\ovi\\ $\\lambda\\lambda 1032,1038$ doublet. Most of these intrinsic absorption systems show multiple components with intrinsic widths of 100 \\kms\\ spread over a blue-shifted velocity range of less than 1000 \\kms. Galaxies in our sample with existing X-ray or longer wavelength UV observations also show \\civ\\ absorption and evidence of a soft X-ray warm absorber. In some cases, a UV absorption component has physical properties similar to the X-ray absorbing gas, but in others there is no clear physical correspondence between the UV and X-ray absorbing components. Models in which a thermally driven wind evaporates material from the obscuring torus naturally produce such inhomogeneous flows. ", "introduction": "Roughly 50\\% of all Seyfert galaxies show UV absorption lines, most commonly seen in {\\sc C~iv} and Ly$\\alpha$ \\citep{Crenshaw99}. X-ray ``warm absorbers\" are equally common in Seyferts \\citep{Reynolds97, George98}. All instances of X-ray absorption also exhibit UV absorption \\citep{Crenshaw99}. While \\citet{Mathur94} and \\citet{Mathur95} have suggested that the same gas gives rise to both the X-ray and UV absorption, the spectral complexity of the UV and X-ray absorbers indicates that a wide range of physical conditions are present. Multiple kinematic components with differing physical conditions are seen in both the UV \\citep{Crenshaw99, Kriss00b} and in the X-ray \\citep{Reynolds97, Kriss96, Kaspi01}. The short wavelength response (912--1187 \\AA) of the {\\it Far Ultraviolet Spectroscopic Explorer (FUSE)} \\citep{Moos00} enables us to make high-resolution spectral measurements ($R \\sim 20,000$) of the high-ionization ion {\\sc O~vi} and the high-order Lyman lines of neutral hydrogen. The \\ovi\\ doublet is a crucial link for establishing a connection between the higher ionization absorption edges seen in the X-ray and the lower ionization absorption lines seen in earlier UV observations. The high-order Lyman lines provide a better constraint on the total neutral hydrogen column density than Ly$\\alpha$ alone. Lower ionization species such as {\\sc C~iii} and {\\sc N~iii} also have strong resonance lines in the \\FUSE\\ band, and these often are useful for setting constraints on the ionization level of any detected absorption. The Lyman and Werner bands of molecular hydrogen also fall in the \\FUSE\\ band, and we have searched for intrinsic $\\rm H_2$ absorption that may be associated with the obscuring torus. We have been conducting a survey of the $\\sim100$ brightest AGN using {\\it FUSE}. As of November 1, 2002, we have observed a total of 87; of these, 57 have $z < 0.15$, so that the {\\sc O~vi} doublet is visible in the FUSE band. ", "conclusions": "The multiple kinematic components frequently seen in the UV absorption spectra of AGN clearly show that the absorbing medium is complex, with separate UV and X-ray dominant zones. In some cases, the UV absorption component corresponding to the X-ray warm absorber can be clearly identified (e.g., Mrk~509) \\citep{Kriss00b}. In others, however, {\\it no} UV absorption component shows physical conditions characteristic of those seen in the X-ray absorber (NGC~3516, NGC~5548) \\citep{Kriss96, Brotherton02}. One potential geometry for this complex absorbing structure is high-density, low-column UV-absorbing clouds embedded in a low-density, high-ionization medium that dominates the X-ray absorption. This is possibly a wind driven off the obscuring torus \\citep{KK95, KK01}. At the critical ionization parameter for evaporation, there is a broad range of temperatures that can coexist in equilibrium at nearly constant pressure; for this reason, the flow is expected to be strongly inhomogeneous. What would this look like in reality? As a nearby analogy, consider the HST images of the pillars of gas in the Eagle Nebula, M16. These show the wealth of detailed structure in gas evaporated from a molecular cloud by the UV radiation of nearby newly formed stars \\citep{Hester96}. Figure 1 shows what this might look like in an AGN---a dense molecular torus surrounded by blobs, wisps, and filaments of gas at various densities. It is plausible that the multiple UV absorption lines seen in AGN with warm absorbers are caused by high-density blobs of gas embedded in a hotter, more tenuous, surrounding medium, which is itself responsible for the X-ray absorption. Higher density blobs would have lower ionization parameters, and their small size would account for the low overall column densities. \\begin{figure}[ht] \\centerline{\\epsfxsize=3.7in\\epsfbox{gkriss_agn2003_fig1.ps}} \\caption{ An artist's conception of how a molecular torus surrounding an AGN might appear based on HST images of the Eagle Nebula. Note the complex of wisps and blobs of gas close to the surface of the molecular material. \\label{torus}} \\end{figure} At sight lines close to the surface of the obscuring torus, one might expect to see some absorption due to molecular hydrogen. Given the dominance of Type 1 AGN in our observations so far, the lack of any intrinsic $\\rm H_2$ absorption is not too surprising since our sight lines are probably far above the obscuring torus. NGC~4151 and NGC~3516 are examples where the inclination may be more favorable since these objects have shown optically thick Lyman limits in the past \\citep{Kriss92, Kriss96}, but our FUSE observations do not show such high levels of neutral hydrogen. Molecular hydrogen will not survive long in an environment with a strong UV flux, and this probably accounts for the lack of $\\rm H_2$ absorption. In summary, we find that \\ovi\\ absorption is common in low-redshift ($z < 0.15$) AGN. 30 of 53 Type 1 AGN with $z < 0.15$ observed using \\FUSE\\ show multiple, blended \\ovi\\ absorption lines with typical widths of $\\sim 100~\\kms$ that are blueshifted over a velocity range of $\\sim$ 1000 \\kms. Those galaxies in our sample with existing X-ray or longer wavelength UV observations also show {\\sc C~iv} absorption and evidence of a soft X-ray warm absorber. In some cases, a UV absorption component has physical properties similar to the X-ray absorbing gas, but in others there is no clear physical correspondence between the UV and X-ray absorbing components." }, "0403/astro-ph0403227_arXiv.txt": { "abstract": "With its excellent spatial resolution, low background, and hard-band response, the {\\em Chandra} X-ray Observatory is ideal for performing exploratory surveys. These efficient, sensitive observations can place constraints on fundamental properties of a quasar continuum including the X-ray luminosity, the ratio of X-ray to UV power, and the X-ray spectral shape. To demonstrate the power of such surveys to provide significant insight, we consider two examples, a Large Bright Quasar Survey sample of broad absorption line quasars and a sample of Sloan Digital Sky Survey (SDSS) quasars with extreme C~{\\sc iv} blueshifts. In both cases, exploratory {\\em Chandra} observations provide important information for a physical understanding of UV spectroscopic differences in quasars. ", "introduction": "X-ray emission appears to be a universal signature of quasar spectral energy distributions, confirming expectations from accretion physics. Based on rapid variability of soft X-rays in conjunction with the standard black-hole paradigm, these photons are believed to be emitted from the region immediately surrounding the black hole. Energetically, X-rays are significant, contributing 2--20$\\%$ of the bolometric luminosity. X-ray observations are thus an important component of any multi-wavelength campaign to probe quasar populations. The excellent spatial resolution of the {\\em Chandra} High Resolution Mirror Assembly and the effective background rejection of the ACIS instrument make this combination uniquely powerful for quasar surveys. For reference, during a 5~ks observation, the 0.5--8.0~keV background within a $2\\arcsec$-radius source region is typically $\\sim$0.1~ct. Because ACIS is photon-limited even beyond $100$~ks (Alexander et al.\\ 2003), the point-source detection limit scales {\\em linearly} with exposure time, unlike the $\\sqrt{t} $ dependence common in other wavelength bands. In conjunction with sub-arcsec positional accuracy, known optical point sources can be robustly detected with 3--5 photons. In 5 ks, this corresponds to a \\mbox{0.5--8.0~keV} flux of $\\sim7\\times10^{-15}$~erg~cm$^{-2}$~s$^{-1}$ for a typical quasar X-ray spectrum. A 3--7~ks {\\em Chandra} exposure, the regime of exploratory observations, is generally insufficient for gathering enough X-rays for spectral analysis of a quasar. However, the strategy of exploratory observations enables the extension of results from spectroscopic observations of individual targets to larger, well-defined samples, and the investigation of connections between X-ray properties and other wavelength regimes. From these datasets, standard X-ray observables are \\mbox{0.5--8.0~keV} flux, hardness ratio,\\footnote{The hardness ratio is defined to be $(h-s)/f$, where $h$=2--8~keV ct, $s$=0.5--2.0~keV ct, and $f$=0.5--8.0~keV ct.} and $\\alpha_{\\rm ox}$.\\footnote{The quantity $\\alpha_{\\rm ox}$ equals $0.384\\log(f_{\\rm X}/f_{\\rm 2500})$ where $f_{\\rm X}$ and $f_{\\rm 2500}$ are the flux densities at rest-frame 2~keV and 2500~\\AA, respectively.} We briefly describe the initial results from two exploratory {\\em Chandra} quasar surveys to illustrate the utility of this observing strategy. Other examples in the literature of successful applications of this approach to quasar studies include surveys of high-$z$ (e.g., Brandt et al.\\ 2002; Vignali et al.\\ 2003), red (Wilkes et al. 2002), and X-ray weak (Risaliti et al.\\ 2003) quasars. \\begin{figure} \\plotone{fig1.eps} \\caption{Hardness ratios versus $\\alpha_{\\rm ox}$ from exploratory {\\em Chandra} surveys for two quasar samples. ($a$) The Large Bright Quasar Survey broad absorption line quasar sample (Gallagher et al.\\ 2003). ($b$) The SDSS C~{\\sc iv} blueshift sample (Richards et al., in prep.).} \\end{figure} \\vspace{-0.1cm} ", "conclusions": "" }, "0403/astro-ph0403233_arXiv.txt": { "abstract": "We present a newly developed cosmological hydrodynamics code based on weighted essentially non-oscillatory (WENO) schemes for hyperbolic conservation laws. WENO is a higher order accurate finite difference scheme designed for problems with piecewise smooth solutions containing discontinuities, and has been successfully applied for problems involving both shocks and complicated smooth solution structures. We couple hydrodynamics based on the WENO scheme with standard Poisson solver - particle-mesh (PM) algorithm for evolving the self-gravitating system. The third order low storage total variation diminishing (TVD) Runge-Kutta scheme has been used for the time integration of the system. To test accuracy and convergence rate of the code, we subject it to a number of typical tests including the Sod shock tube in multidimensions, the Sedov blast wave and formation of the Zeldovich pancake. These tests validate the WENO hydrodynamics with fast convergence rate and high accuracy. We also evolve a low density flat cosmological model ($\\Lambda$CDM) to explore the validity of the code in practical simulations. ", "introduction": "Though the universe seems to be dominated by the dark sides of both matter and energy (Turner, 2002), the observed luminous universe has been existing in the form of baryonic matter, whose mass density, constrained by the primordial nucleosynthesis (Walker, et al., 1991), only occupies a small amount of the total density. To account for the observational features revealed by the baryonic matter, i.e., X-ray emitting gas in galaxies and clusters (Mulchaey, 2000), intergalactic medium inferred from Ly$\\alpha$ forest (Rauch, 1998), X-ray background radiation (Giacconi et al. 1962) and distorted spectrum of the cosmic background radiation due to the Sunyaev-Zeldovich effect (Zel'dovich \\& Sunyaev 1969; Ostriker \\& Vishniac 1986) etc., it would be necessary to incorporate hydrodynamics into cosmological investigations. This motivation has stimulated great efforts to apply a variety of gas dynamics algorithms to cosmological simulations. For a general review of the state-of-the-art on this topic, we refer to Bertschinger (1998). Due to the high non-linearity of gravitational clustering in the universe, there are two significant features emerging in cosmological hydrodynamic flow, which pose more challenges than the typical hydrodynamic simulation without self-gravity. One significant feature is the extremely supersonic motion around the density peaks developed by gravitational instability, which leads to strong shock discontinuities within complex smooth structures. Another feature is the appearance of an enormous dynamic range in space and time as well as in the related gas quantities. For instance, the hierarchical structures in the galaxy distribution span a wide range of length scales from a few kpc resolved by individual galaxy to several tens of Mpc characterizing the largest coherent scale in the universe. A variety of numerical schemes for solving the coupled system of collisional baryonic matter and collisionless dark matter have been developed in the past decades. They fall into two categories, particle methods and grid based methods. The particle methods include variants of the smooth-particle hydrodynamics (SPH; Gingold \\& Monagham 1977, Lucy 1977) such as those of Evrard (1988), Hernquist \\& Katz (1989), Navarro \\& White (1993), Couchman, Thomas \\& Pierce (Hydra, 1995), Steinmetz (1996), Owen et al. (1998) and Springel, Yoshida \\& White (Gadget, 2001). The SPH method solves the Lagrangian form of the Euler equations, and could achieve good spatial resolutions in high density regions, but works poorly in low density regions. It also suffers from degraded resolution in shocked regions due to the introduction of sizable artificial viscosity. The grid based methods are to solve the Euler equations on structured or unstructured grids. The early attempt was made by Cen (1992) using a central difference scheme. It uses artificial viscosity to handle shocks and has first-order accuracy. The modern approaches implemented for high resolution shock capturing are usually based on the Godunov algorithm. The two typical examples are the total-variation diminishing (TVD) scheme (Harten 1983) and the piecewise parabolic method (PPM) (Collella \\& Woodward 1984). Both schemes start from the integral form of conservation laws of Euler equations and compute the flux vector based on cell averages (finite volume scheme). The TVD scheme modifies the flux using an approximate solution of the Riemann problem with corrections added to ensure that there are no postshock oscillations. While in the PPM scheme, the Riemann problem is solved accurately using a quadratic interpolation of the cell-average densities that is constrained to minimize postshock oscillations. In the cosmological setting, the TVD based codes include those of Ryu et al. (1993), the moving-mesh scheme (Pen, 1998), and the smooth Lagrangian method (Gnedin, 1995); and the PPM based codes include those of Stone \\& Norman (Zeus; 1992), Bryan et al. (1995), Sornborger et al. (1996), Ricker, Dodelson \\& Lamb (COSMOS; 2000). The grid-based methods suffer from the limited spatial resolution, but they work extremely well both in low and high density regions as well as in shocks. To reach a large dynamical range, the Godunov methods have also been implemented with adaptive mesh refinement (RAMSES: Teyssier, 2002; ENZO: Norman \\& Bryan, 1999; O'Shea et al., 2004), which is more adequate to explore the fine structures in the hydrodynamic simulation. We describe in this paper an alternative hydrodynamic solver which discretizes the convection terms in the Euler equations by the fifth order finite difference WENO (weighted essentially non-oscillatory) method, first developed in Jiang \\& Shu (1996), with a low storage third order Runge-Kutta time discretization, which was proven to be nonlinearly stable in Gottlieb \\& Shu (1998). The WENO schemes are based on the essentially non-oscillatory (ENO) schemes first developed by Harten et al. (1987) in the form of finite volume scheme for hyperbolic conservative laws. The ENO scheme generalizes the total variation diminishing (TVD) scheme of Harten (1983). The TVD schemes typically degenerate to first-order accuracy at locations with smooth extrema while the ENO scheme maintains high order accuracy there even in multi-dimensions. WENO schemes further improve upon ENO schemes in robustness and accuracy. Both ENO and WENO schemes use the idea of adaptive stencils in the reconstruction procedure based on the local smoothness of the numerical solution to automatically achieve high order accuracy and non-oscillatory property near discontinuities. For WENO schemes, this is achieved by using a convex combination of a few candidate stencils, each being assigned a nonlinear weight which depends on the local smoothness of the numerical solution based on that stencil. WENO schemes can simultaneously provide a high order resolution for the smooth part of the solution, and a sharp, monotone shock or contact discontinuity transition. WENO schemes are extremely robust and stable for solutions containing strong shocks and complex solution structures. Moreover, a significant advantage of WENO is its ability to have high accuracy on coarser meshes and to achieve better resolution on the largest meshes allowed by available computer memory. We will describe the fifth order WENO scheme employed in this paper briefly in \\S 3. For more details, we refer to Jiang \\& Shu (1996) and the lecture notes by Shu (1998, 1999). WENO schemes have been widely used in applications. Some of the examples include dynamical response of a stellar atmosphere to pressure perturbations (Zanna, Velli \\& Londrillo, 1998); shock vortex interactions and other gas dynamics problems (Grasso \\& Pirozzoli, 2000a; 2000b); incompressible flow problems (Yang et al., 1998); Hamilton-Jacobi equations (Jiang \\& Peng, 2000); magneto-hydrodynamics (Jiang \\& Wu, 1999); underwater blast-wave focusing (Liang \\& Chen, 1999); the composite schemes and shallow water equations (Liska \\& Wendroff, 1998, 1999); real gas computations (Montarnal \\& Shu, 1999), wave propagation using Fey's method of transport (Noelle, 2000); etc. In the context of cosmological applications, we have developed a hybrid N-body/hydrodynamical code that incorporates a Lagrangian particle-mesh algorithm to evolve the collisionless matter with the fifth order WENO scheme to solve the equations of gas dynamics. This paper is to detail this code and assess its accuracy using some numerical tests. We proceed as follows. In \\S 2, we present the basic cosmological hydrodynamic equation for the baryon-CDM coupling system. \\S 3 gives a brief discussion of the numerical scheme for solving the hydrodynamic equations, especially about the implementation of the finite difference fifth order WENO scheme and the TVD time discretization. In \\S 4, we validate the code using a few challenging numerical tests. Concluding remarks are drawn in \\S 5. ", "conclusions": "In this paper, we have described a newly developed hybrid cosmological hydrodynamic code based on weighted essentially non-oscillatory (WENO) schemes for the Euler system of conservation laws. We implement the fifth order finite difference WENO to solve the inviscid fluid dynamics on a uniform Eulerian grid combining with a third order low storage Runge-Kutta TVD scheme for advancing in time. In order to solve the cosmological problem involving both collisional baryonic matter and collisionless dark matter, we incorporate the particle-mesh method for computing the self-gravity into our cosmological code. The code has been subjected to a number of tests for its accuracy and convergence. As expected, the WENO scheme demonstrates its capacity of capturing shocks and producing sharp and non-oscillatory discontinuity transition without generating oscillations. In comparison with other existing hydrodynamic codes such as the TVD or PPM schemes, one striking feature of the WENO code is that it retains higher order accuracy in smooth regions including at smooth extrema even in multidimensions, and yet it is still highly stable and robust for strong shocks. In performance, the WENO scheme needs more floating point operations per cell than those of the PPM and TVD schemes. However, in compensating for twice or more loss of the computational speed, the WENO scheme achieves both higher order accuracy and convergence rate than PPM and TVD codes according to our numerical tests. In the presence of gravity, the hydrodynamics become more challenging than that without gravity due to the highly non-linearity of gravitational clustering. One serious problem encountered in many cosmological applications is the so called high Mach number problem. To address this problem, we have incorporated an extra technique into our cosmological WENO/PM code, which is actually a combination of the dual energy algorithm (Bryan et al. 1995) and the energy-entropy algorithm (Ryu et al. 1993). Namely, instead of solving the internal energy equation in regions free of shocks as was done in the dual energy algorithm (Bryan et al. 1995), we solve the modified entropy equation (Ryu et al. 1993), which takes a conservative form and can be easily solved using the standard WENO scheme. This improvement over our hydrodynamic WENO code ensures an accurate tracking of the temperature field in regions free of shocks. It is pointed out that the high order WENO discretization, e.g., the fifth order WENO scheme adopted in this paper, introduces a quite small numerical viscosity, which does not lead to a significant violation of energy conservation in the presence of gravitational fields. While for a second-order TVD scheme, the numerical diffusion is no longer negligible. In order to have a better conservation of the total energy, it is usually corrected by adding a compensation term in the gravitational force term (Ryu et al, 1993). The Euler hydrodynamics on fixed meshes have several distinct advantages which includes simplicity for implementation, easy data parallelization, relatively low floating point cost, large dynamic range in mass and high resolution of shock capturing. In particular, the WENO scheme can also achieve a higher accuracy on coarse meshes and a better resolution on the largest meshes allowed by available memory. To suit for large simulations of the cosmological problem, further improvement of the hydrodynamic WENO code is needed in its implementation on distributed memory computers. The parallel version of the code based on the Message-Passing Interface (MPI) has been under development." }, "0403/astro-ph0403005_arXiv.txt": { "abstract": "This is the written version of an invited review talk for the 13 Feb 2004 AAAS Meeting in Seattle. The talk's goal is to present a philosophical view of extragalactic astronomy as it applies to the sub-field of galaxy evolution. The talk is divided into three parts: 1) How we got to where we are (technology drivers to our science goals), 2) What's new and special (how that technology has achieved our recent science results) and 3) How an improved worldview will help us in the near future. The intended audience for this talk is a generally knowledgeable scientist, but not an astronomer by training. This talk is also *not* intended to be a complete review of the field of galaxy evolution and only includes a few recent results extracted from the astro-ph archives to present the current state of our field. ", "introduction": "Historians will no doubt look back onto the last part of the 20th century as a true \"Golden Age\" for observational cosmology and galaxy evolution due to the enormous advancements driven primarily by technological leaps in the telescopes, detectors and computers. Like most modern sciences, the study of galaxies is beyond the basic human senses (although the Magellenic Clouds are visible to the naked eye in the southern hemisphere and Andromeda is at the limit of the human eye for very dark sites). Thus, the study of galaxies did not even begin until the development of 1m class telescopes at the end of the 19th century. With the discovery of an expanding Universe (Hubble 1936) and the confirmation of a Creation point (Penzias \\& Wilson 1965), it was quickly realized that astronomers have a unique capability, that no other science has. This is the capability to actually follow the evolution of distant objects while the size of the Universe is much greater than the speed of light (i.e. lookback time). To imagine the impact of lookback time on a field of science, imagine the response of a palaeontologist if offered a device that would allow them to observe animal behavior during the Jurassic era, or a musicologist who is allowed to listen in on Mozart's practice sessions. Lookback time means that the more distant an object, the farther into our past it is. Thus began in extragalactic astronomy the great hunt for high redshift objects, redshift being a measure of distance in an expanding Universe. Of course, more distant also means fainter in apparent luminosity, which drives the need for larger and larger telescopes to collect the photons from high redshift, or very distant, galaxies. Our first quantum jump in the study of galaxy evolution was during the 1950's and 60's with the development of the Palomar 5m and NOAO's 4m telescopes. With this technology, spectroscopy out to several billion light-years was doable. Solid state technology in the 80's gave us highly efficient and highly reliable (for photometry) detectors. The 90's gave us 10m class telescopes and the Hubble Space Telescope, arguably the most important instrument for the study of galaxy evolution. Parallel to observational achievements were breakthroughs in many theoretical arenas, primarily the coming of age of the computer simulation in the 80's as the main tool for astrophysical research (observers call themselves astronomers, while theorists call themselves astrophysicists for some strange reason). During the 90's, decreasing access to telescopes (due to an increasing population of observers) and decreasing computer hardware costs produced a glut of theoretical Ph.D. theses. Within a decade, our field went from the point where theorists numbered one for every four observers to near parity. When combined with uncertain values for the basic cosmological constants (i.e. $H_o$ and $q_o$), this lead to an explosion of ideas, but with very few constraints to narrow our focus on what was really happening after galaxy formation. Unfortunately, the same technology that allows us to observe the past also confines our window of knowledge. For example, galaxies present themselves over a range of luminosities, which corresponds directly with the mass of the galaxy (i.e. the number of stars). While the brightest galaxies (up to 10$^{13}$ $M_{\\sun}$) are the most dominant members of clusters of galaxies, and the easiest to obtain photometric and spectroscopic information, they are not the dominant members of the Universe by number. That honor goes to the small, dwarf galaxies (masses around 10$^8$ $M_{\\sun}$) who number over a factor of 10 to 100 more numerous than the brightest galaxies in clusters. Thus, the limiting magnitude of your telescope (the depth in luminosity that can be achieved for a decent S/N) will allow you to study bright galaxies to distant redshifts (far into the past) or both bright and faint galaxies at lower redshifts. This bias forced our conclusions about galaxy evolution in the 80's and 90's to be limited to only the very brightest galaxies in rich environments. The newest technology allows us to address the full galaxy population at cosmologically interesting distances of 0.5 to 0.7 (see Figure 1.1). \\begin{figure} \\centering \\includegraphics[width=12cm]{lum_morph.epsf} \\caption{ This figure displays the effects of limiting magnitude to the ability of galaxy evolution studies to resolve their science goals. Early technology (shown by the 5m limit) shows that only the brightest galaxies could be measured at cosmologically interesting distances. The Hubble Deep Field (HDF) extended the depth to more normal galaxies and the Hubble Ultra Deep Field (HUDF) will allow a comparison of the evolution of dwarf and giant galaxies, a key test to the hierarchical merger scenario of galaxy formation. } \\end{figure} The greatest progress with respect to the study of galaxy evolution was the determination of the cosmic distance scale parameters, $H_o$ and $q_o$ by the Hubble Cepheid Key Project and the High Redshift Supernova Cosmology Project. While these two parameters are not directly related to galaxy evolution parameters, their determination has finally released researchers from being forced to interpret evolutionary effects through a myriad of combinations of $H_o$ and $q_o$. Or, worse, attempting to deduce these cosmological parameters from poorly known evolutionary effects. Galaxy evolutionists are now free to focus solely on the star formation history of galaxies with lookback time determined solely by the redshift of the galaxy. ", "conclusions": "" }, "0403/astro-ph0403696_arXiv.txt": { "abstract": "We revisit the notion that galaxy motions can efficiently heat intergalactic gas in the central regions of clusters through dynamical friction. For plausible values of the galaxy mass-to-light ratio, the heating rate is comparable to the cooling rate due to X-ray emission. Heating occurs only for supersonic galaxy motions, so the mechanism is self-regulating: it becomes efficient only when the gas sound speed is smaller than the galaxy velocity dispersion. We illustrate with the Perseus cluster, assuming a stellar mass-to-light ratio for galaxies in the very central region with the dark-matter contribution becoming comparable to this at some radius $r_s$. For $r_s \\la 400~{\\rm kpc} \\sim 3 r_{\\rm cool}$---corresponding to an average mass-to-light ratio of $\\sim10$ inside that radius---the dynamical-friction coupling is strong enough to provide the required rate of gas heating. The measured sound speed is smaller than the galaxy velocity dispersion, as required by this mechanism. With this smaller gas temperature and the observed distribution of galaxies and gas, the energy reservoir in galactic motions is sufficient to sustain the required heating rate for the lifetime of the cluster. The galaxies also lose a smaller amount of energy through dynamical friction to the dark matter implying that non--cooling-flow clusters should have flat-cored dark-matter density distributions. ", "introduction": "\\label{sec:intro} Galaxy cluster gas loses thermal energy copiously through X-ray emission. In the absence of energy input, radiative cooling in cores of clusters should result in substantial gas inflow (see Fabian 1994 for a review). These ``cooling flows'' would have associated mass deposition rates of several hundred ${\\rm M_\\odot yr^{-1}}$ in some clusters (Peres \\etal 1998). Nevertheless, recent high-resolution X-ray observations (e.g., Peterson \\etal 2001, 2003) have revealed that there is little evidence for the expected multi-phase gaseous structures, strongly suggesting that mass dropout is being prevented by some source that is heating the gas, thereby balancing radiative energy loss in the central region of clusters. Recent work has focused on two prospective heating mechanisms: (1) diffusive heat transport, via thermal conduction (\\pcite{tuc83,bre88,nar01,voi02,fab02,zak03,kim03a}) and/or turbulent mixing (\\pcite{cho03,kim03b,voi04}), from hotter gas in the outer region to that in the core; (2) energy input from jets, outflows, and radiation from a central active galactic nucleus (\\pcite{cio01,chu02,bru02,kai03}). There is however at least one additional mechanism that appears to have been overlooked. It involves the energy lost by concentrated clumps of matter (galaxies) as they move through the cluster. Part of this energy may go into re-arranging the dark-matter mass distribution (El-Zant \\etal 2003), but a significant fraction should end up deposited in the gas. That dynamical-friction (DF) coupling can transform the dynamical energy of galaxies into thermal motion in the gas has been known for at least four decades (e.g., Dokuchaev 1964). Relatively recent work involving this notion includes the analysis by Miller (1986) of the Perseus cluster and that of Just \\etal (1990) concerning the Coma cluster. Both studies confirm that, provided that the mass-to-light ratio of galaxies is $\\sim 20$, energy lost by galaxies to the gas should be sufficient in counteracting the radiative cooling of the gas in the central region of these clusters. Several developments on the empirical side suggest renewed relevancy for this mechanism. One obvious one involves recent X-ray data confirming that a heating mechanism {\\em is} required; whereas the consensus in the 1980's was against this conclusion, it now seems inescapable. The second involves revisions to the inferred gas electron densities in the central region of clusters; best values referred to by Miller and Just \\etal are of the order of $10^{-3} {\\rm cm^{-3}}$, while current best estimates are rather in the range of $10^{-2}-10^{-1} {\\rm cm^{-3}}$ (Kaastra \\etal 2004). This leads to an order-of-magnitude increase in the dynamical friction coupling between the galaxies and gas. There has also been progress in determining the mass-to-light ratio of galaxies. On the theoretical side, work by Just \\etal (1990) and Ostriker (1999) has since shed some light on the behavior of the dynamical-friction coupling in a gaseous medium in the transonic and subsonic regimes. There appears to be {\\em a priori} no reason why the rate of energy loss from galaxies to gas via dynamical friction should be of the same order of that radiated by the gas. However, {\\em this coupling is active only when the sound speed of the gas is smaller than the typical velocity of galaxies}. The mechanism is therefore self-regulating; the gas cools until the dynamical-friction heating rate is always equal to the cooling rate. We start by pointing out, in the next Section, why this is expected to be so, before moving on to develop a Monte-Carlo model to estimate the total energy expected to flow into the cooling region of the Perseus cluster, using recent data for galaxy luminosities and projected positions, as well as for the gas parameters of that cluster, and averaging over a set of different realizations where three-dimensional positions, velocities, and mass-to-light ratios are treated as stochastic quantities. The final Section discusses briefly the central dark-matter distribution and energetics, as well as some remaining questions. ", "conclusions": "\\label{sec:disc} That dynamical-friction coupling between cluster galaxies and gas can significantly heat the latter component is not a new concept. Indeed Miller (1986) has pointed out that a few luminous galaxies in the Perseus cluster's central region may alone deposit sufficient energy to keep gas in that region from cooling, provided that they had a mass-to-light ratio of about 20. We revisited this cluster using recently compiled galaxy and gas data and determinations of the mass-to-light ratios in galaxies. Our Monte Carlo model shows that the rate at which energy is deposited by galaxies into the gas within the cooling radius can completely compensate for radiative loss from within that radius, if galaxies at radii $\\la 400$ kpc from the center of the cluster have dark mass that is comparable to their stellar mass (with galaxies at smaller radii having progressively smaller dark-matter content, vanishing for those near the center of the cluster). The associated average mass-to-light ratio within this radius is about 10. A robust and potentially important prediction of the model is that the drop in gas temperature invariably observed in the cooling regions of clusters is necessary for the DF coupling to be sufficiently strong. There is a natural interpretation for this phenomenon: the gas cools until dynamical-friction heating becomes sufficiently efficient to prevent a further drop in temperature. The gas random motion is then coupled to that of the galaxies, and so remains in approximate equilibrium with it. All this simply follows from the fact that dynamical friction drops sharply for subsonic motion. In the case of the Perseus cluster, the transition from weak to strong coupling occurs precisely in the range of sound speeds found in the cooling core of the cluster. The action of the mechanism discussed in this paper also has consequences for the dark-matter distribution. As shown by El-Zant \\etal (2003), dynamical friction from the galaxies will heat a density cusp, creating a core with radius corresponding to roughly a fifth of the original Navarro, Frenk \\& White (1997) scale length (cf., Fig.~1 of El-Zant et al.), where dark matter has been spread out to larger radii (Fig.~3 of El-Zant et al.). Approximating the initial density distribution in this region such that $\\rho_i = \\rho_0 (r_0/r)$ and the final one with $\\rho_f = \\rho_0$, the energy required for this transformation is $\\Delta \\Phi = (8 \\pi G)^{-1} \\int_{0}^{r_0} (|d \\phi/dr|_i^2 - |d\\phi/dr|_f^2) 4 \\pi r^2 dr = 22 \\pi^2 \\rho_0^2 r_0^5/45$ [with $d\\phi/dr = (4 \\pi G/r^2 )\\int \\rho r^2 dr$]. Taking $r_0 = r_{\\rm cool}$ and $\\rho_0$ to correspond to four times the value defined by the electron density $\\rho_e (r_0) = 0.0053$ (where we have used the gas to gravitational mass ratio in Table~5 of Peres \\etal 1998 and equation (4) of Churazov \\etal 2003 for the electron density), one finds $\\Delta \\Phi = - 6 \\times 10^{60}~{\\rm erg}$, which is comparable to the binding energy of the CD galaxy $\\Phi_{\\rm CD} \\approx - (G M_{\\rm CD}^2/R_{\\rm CD})$. It is however significantly smaller than the energy radiated from gas inside the cooling radius for the age of the cluster (e.g., for five Gyr this amounts to $8 \\times 10^{61}\\, {\\rm erg}$). Nevertheless, this energy is easily available from fast moving galaxies from beyond the very inner region---for example its value coincides with the kinetic energy in a mass similar to that of the CD galaxy and moving at $2000~\\kms$, and that material can be supplied solely by the stellar mass of galaxies within $2 r_{\\rm cool}$---the region from which, according to our model, the bulk of energy input to the gas comes from. It is worth noting here that, under the circumstances just described, the mass of the gas within $r_{\\rm cool}$ [which using equation (4) of Churazov \\etal 2003, adjusted for $h=0.7$ with $r_{\\rm cool} = 130\\,{\\rm kpc}$ amounts to $2.2 \\times 10^{12} {\\rm M_\\odot}$] is comparable to that of the galaxies in that region. Due to the temperature drop in the gas, thermal motions can be significantly smaller than that of the galaxies. The energy of the gas can therefore be smaller than that of the galaxies by up to an order of magnitude. There is therefore sufficient energy in galaxies to heat the gas many times over, with the crucial parameter, which was the focus of this paper, being the rate at which this is transferred. Furthermore, once the dark matter in the center has been heated, it absorbs little additional energy from the galaxies, since dynamical friction does not act on fast particles, and the coupling with the dark matter at larger radii decreases rapidly with radius (see El-Zant \\etal 2003 for further discussion). The energy lost to the dark matter is therefore less than that going into keeping the gas at constant temperature. The mass of the CD galaxy, as well as the spatial and velocity distribution of galaxies in the central region of the cluster will reflect the history of energy loss to both components. Several questions remain open. Prominent among these is the issue of thermal stability. From the condition that thermal stability requires a remarkably narrow range of heating rates, Bregman \\& David (1989) have argued against Miller's proposition that the Perseus cluster gas is heated by DF from galaxies. However the functional form of the heating rate used by these authors does not agree with later calculations that have been borne out by numerical studies---e.g., their postulated form has a {\\em negative} energy transfer rate ($\\sim -1/V^3$) for highly subsonic velocities, resulting in gas {\\em cooling}, which is incompatible with the positive (even if small) heating found by Just \\etal (1990) and Ostriker (1999) in that limit. It will be therefore necessary to repeat these calculations using adequate forms for the heat-loss function. These calculations would also address another crucial question, that concerning the precise manner in which the energy lost by the galaxies is distributed in the gas. We have adopted the simple approach where this energy is deposited isotropically and equally in logarithmic intervals. This should approximate how the energy is initially deposited, at least for highly supersonic galaxies (and since, in the cooling region, the sound speed may be several times smaller than the velocity dispersion, this may not be a very bad assumption). Even then, however, the deposited energy may still be transported toward the center of the cluster by wave motions, as suggested, for example, by Balbus \\& Soker (1990), and redistributed. We note here that the claim of these authors that DF from galaxies is insufficient to heat the cluster core is based upon outdated values for the electron density and an outdated ($1/r$) gas distribution. With updated values for the electron density and gas distribution (e.g., the empirical formula of Churazov \\etal 2003), their conclusions are changed, as we have shown. We thus conclude that it is far from obvious that DF-heating of cluster gas is irrelevant." }, "0403/astro-ph0403375_arXiv.txt": { "abstract": "{ We present the \\XMM{} spectra of three low-redshift intermediate Seyferts (one Sy1.5, and two Sy1.8), from our survey of hard spectrum \\Rosat{} sources. The three AGN are well fitted by absorbed powerlaws, with intrinsic nuclear photoelectric absorption from column densities between 1.3 and 4.0 $\\times 10^{21}$~cm$^{-2}$. In the brightest object the X--ray spectrum is good enough to show that the absorber is not significantly ionized. For all three objects the powerlaw slopes appear to be somewhat flatter ($\\Gamma\\sim1.3-1.6$) than those found in typical unabsorbed Seyferts. The constraints from optical and X--ray emission lines imply that all three objects are Compton-thin. For the two fainter objects, the reddening deduced from the optical broad emission lines in one of them, and the optical continuum in the other, are similar to those expected from the X--ray absorption, if we assume a Galactic gas-to-dust ratio and reddening curve. The broad line region Balmer decrement of our brightest object is larger than expected from its X--ray absorption, which can be explained either by an intrinsic Balmer decrement with standard gas-to-dust ratio, or by a $>$Galactic gas-to-dust ratio. These $\\ge$~Galactic ratios of extinction to photoelectric absorption cannot extend to the high redshift, high luminosity, broad line AGN in our sample, because they have column densities $>10^{22}$~cm$^{-2}$, and so their broad line regions would be totally obscured. This means that some effect (e.g., luminosity dependence, or evolution) needs to be present in order to explain the whole population of absorbed AGN. ", "introduction": "According to the unified model for Active Galactic Nuclei (AGN)(Antonucci 1993), broad-line Seyferts (type 1) and narrow-line (type 2) Seyferts are intrinsically the same type of object but are viewed with different orientations to our line of sight. In this model the central engine of the AGN, and the high velocity clouds that produce the broad optical and UV emission lines, are surrounded by a thick torus of dust and cool, molecular gas. In type 1 objects we have a direct view of the central engine and the broadline clouds, whereas in type 2 objects the torus blocks our line of sight to these regions. Such a torus has a large photoelectric opacity at soft X--ray energies, which explains why type 1 Seyferts have steep X--ray spectra with little absorption (Nandra \\& Pounds 1994), and type 2 Seyferts have absorbed X--ray spectra (Smith \\& Done 1996). Narrow emission lines are produced in more distant gas clouds on scales larger than the dusty torus, and so can be seen in both types of object. Objects in which the broad lines are attenuated, but not completely obscured, are called intermediate Seyferts, and are assigned classifications ranging from 1.5-1.9. However, even in the type 1 objects, an extra absorption component from ionized gas is often seen in the X--ray band (George et al. 1998). This ionized gas, the ``warm absorber'', appears to be distributed throughout the BLR and NLR of Seyfert galaxies, and in type 2 objects it can scatter nuclear radiation into our line of sight. The warm absorber has a much lower opacity to soft X--rays than cold gas, but if it contains a significant element of dust (e.g. Brandt, Fabian \\& Pounds 1996) it could produce considerable extinction at optical and ultraviolet wavelengths. An understanding of absorption in AGN is extremely important. For example, unified AGN models for the X--ray background (XRB) (see e.g., Setti \\& Woltjer 1989, for an early proposal, and Gilli, Salvati \\& Hasinger 2001, for a late development) explain the spectrum of the XRB by the superposition of spectra of AGN with various degrees of absorption. Under this scheme, soft spectrum X--ray sources should be mostly type 1 AGN, while hard spectrum X--ray sources would be predominantly type 2 AGN. In the 1990s this matched the observations quite well, because \\Rosat{} surveys in the soft band were dominated by type 1 AGN, (e.g. Mason et al. 2000, Lehman et al. 2001) while surveys selected in harder bands with {\\sl BeppoSAX} (Fiore et al. 1999) were much richer in type 2 AGN. Despite these early successes, recent developments have put in jeopardy the identity between optical type 1 and X--ray unabsorbed objects, on one hand, and optical type 2 and X--ray absorbed objects, on the other. For example, identifications of our survey of \\Rosat{} sources with hard spectra (Page, Mittaz \\& Carrera 2000, 2001) produced mostly type 1 AGN, contrary to the expectations of the unified model. Other examples of X--ray absorbed type 1 objects have been found by Akiyama et al. (2000) using {\\sl ASCA} data, and Mainieri et al. (2002) and Page et al. (2003) using \\XMM{} data. In principle, high gas-to-dust ratios in the X--ray absorbing gas (perhaps due to dust sublimation close to the central X--ray source, Granato, Danese \\& Franceschini 1997), or large dust grains (Maiolino, Marconi \\& Oliva 2001), could give rise to high levels of X--ray absorption, without much optical obscuration. In the opposite sense, Pappa et al. (2001) have found several examples of type 2 AGN with very little or no X--ray absorption. Panessa \\& Bassani (2002) estimate that 10-30\\% of Seyfert 2 galaxies have this property. One striking example is H1320+551 (Barcons, Carrera \\& Ceballos 2003), a Seyfert 1.8/1.9 galaxy with strong optical (BLR and NLR) obscuration, but without any corresponding X--ray absorption from cold gas. The high quality of their \\XMM{} data allows these authors to rule out a warm absorber in this source, which leads them to the conclusion that the BLR is intrinsically reddened in this object: its Sy 1.8/1.9 appearance cannot arise from obscuration of a Seyfert 1 spectrum. The situation is therefore complex. Possible explanations include Compton thick obscuration which could suppress completely the nuclear emission below 10~keV. This spectral range could then be filled by X--rays scattered off the warm absorber, or by extranuclear emission, which would not have in principle a particularly hard or absorbed spectrum. This could result in optically obscured type 2 AGN which appear to be absorption free at X--ray energies. Such a model can in principle be tested, since Bassani et al. (1999) have developed a diagnostic diagram that permits identifying Compton thick sources, as those with high equivalent width Fe emission lines (originating in fluorescence in the torus material), and low 2-10~keV to [OIII] flux ratio. This is based on the observation that [OIII] originates in the NLR, outside the torus, and thus in principle [OIII] should be free of obscuration. Neither H1320+551, nor any of the sources discussed in Pappa et al. (2001), lie in the Compton thick region of this diagram. They represent therefore genuine mismatches between optical and X--ray classifications, at odds with the unified AGN model. Here, we analyze optical and \\XMM{} spectroscopic data on three AGN (RXJ133152.51\\-+111643.5, RXJ163054.25\\-+781105.1, and RXJ213807.61\\--423614.3) from the sample of Page, Mittaz \\& Carrera (2001). Of the objects in this sample which show broad optical emission lines, these three had the highest X--ray fluxes. All three show strong signs of absorption in their \\Rosat{} spectra. In section \\ref{OSp} we present their optical spectra, finding evidence for optical obscuration in at least two of them. We then analyze their \\XMM{} spectra in section \\ref{XSp}, and in particular we measure their intrinsic X--ray absorption. The differences between the levels of optical obscuration and X--ray absorption are discussed in section \\ref{discussion}, as well as a comparison of the X--ray spectral properties of our sources with respect to those of other samples at similar flux levels. Finally, in section \\ref{discussion} we summarize our results. For brevity, we will refer to the three sources using truncated versions of their names (i.e. RXJ1331, RXJ1630 and RXJ2138) in the text. We have used the currently fashionable values of $H_0=70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{m}=0.3$, and $\\Omega_{\\Lambda}=0.7$, throughout this paper. ", "conclusions": "\\label{conclusions} We have presented X--ray spectra from \\XMM{} of the three brightest AGN exhibiting broad optical emission lines (all three are intermediate Seyferts) from the sample of hard spectrum \\Rosat{} sources (Page, Mittaz \\& Carrera 2000, 2001). The X--ray spectra of all three sources are well fitted by powerlaws ($\\Gamma\\sim1.5$), absorbed by moderate amounts of intrinsic nuclear cold material ($N_{\\rm H}\\sim$a few $\\times10^{21}$ cm$^{-2}$). Similarly absorbed sources are an important part of the $S_{\\rm 2-10\\,keV}\\geq 10^{-13}$~cgs AGN population, although the three sources studied here appear to have harder intrinsic power law slopes than the majority of AGN at this flux level. The equivalent width of the narrow emission line at about 6.4~keV found in RXJ1331 is typical of other radio-quiet Compton-thin Seyferts. Detailed analysis of our optical spectroscopic data confirm the classification of these sources as intermediate-type Seyfert galaxies (Sy1.5-Sy1.9). For the two objects in which both H$\\alpha$ and H$\\beta$ are visible (RXJ1331 and RXJ2138), the NLR is shown to be almost free of reddening, while the BLR is significantly reddened. The column density necessary to produce this effect is about 1.5 times that inferred from the X--ray absorption in RXJ1331, and of the order of it in RXJ2138 and RXJ1630 (in this last case from its broad band optical spectrum), if standard Galactic gas-to-dust ratios are assumed. None of the three sources are Compton-thick. The X--ray data for RXJ1331 require that the absorber is cold, allowing us to rule out the presence of dust embedded in a warm absorber in this source. These three low redshift broad line AGN from our sample of hard sources show a ratio of optical extinction to X--ray absorption which is similar to, or larger than, the interstellar medium of our own Galaxy. This cannot be the case for the high luminosity, high redshift, broad line AGN in our sample, because the broad lines would be completely obscured at the X--ray column densities ($>10^{22}$~cm$^{-2}$) observed in these sources. To explain the whole population of absorbed AGN, their effective gas-to-dust ratio must show a large variety, perhaps depending on luminosity, evolving with redshift, or showing geometries different from those proposed by the unified AGN model." }, "0403/astro-ph0403469_arXiv.txt": { "abstract": "Stars form by gravoturbulent fragmentation of interstellar gas clouds. The supersonic turbulence ubiquitously observed in Galactic molecular gas generates strong density fluctuations with gravity taking over in the densest and most massive regions. Collapse sets in to build up stars and star clusters. Turbulence plays a dual role. On global scales it provides support, while at the same time it can promote local collapse. Stellar birth is thus intimately linked to the dynamical behavior of parental gas cloud, which determines when and where protostellar cores form, and how they contract and grow in mass via accretion from the surrounding cloud material to build up stars. Slow, inefficient, isolated star formation is a hallmark of turbulent support, whereas fast, efficient, clustered star formation occurs in its absence. The fact that Galactic molecular clouds are highly filamentary can be explained by a combination of compressional flows and shear. The dynamical evolution of nascent star clusters is very complex. This strongly influences the stellar mass spectrum. The equation of state (EOS) plays a pivotal role in the fragmentation process. Under typical cloud conditions, massive stars form as part of dense clusters. However, for gas with effective polytropic index greater than unity star formation becomes biased towards isolated massive stars, which may be of relevance for understanding Pop III stars. ", "introduction": "\\label{sec:intro} Star clusters form by gravoturbulent fragmentation in interstellar clouds. The supersonic turbulence ubiquitously observed in Galactic gas clouds generates strong density fluctuations with gravity taking over in the densest and most massive regions. Once such cloud regions become gravitationally unstable, collapse sets in and leads to the formation of stars and star clusters. Yet the conditions for fragmentation and the physical processes that govern the early evolution of nascent star clusters are poorly understood. Following up on analytical studies (starting with Jeans 1902; and later Larson 1969; Shu 1977; Elmegreen 1993; Padoan 1995; Padoan \\& Nordlund 2002), most current investigations concentrate on a numerical approach to star cluster formation. For example, the effects of interstellar turbulence have been studied extensively in a series of 3D simulations by Klessen, Burkert, \\& Bate (1998), Klessen, Heitsch, \\& Mac Low (2000), Klessen \\& Burkert (2000, 2001), Heitsch, Mac Low, \\& Klessen (2001a,b), Klessen (2001). See also Ballesteros-Paredes et al.\\ (1999ab, 2003), Padoan \\& Nordlund (1999), Padoan et al.\\ (2001), Bate, Bonnell, \\& Bromm (2003) or Bonnell, Bate, \\& Vine (2003). A complete overview is given in the reviews by Larson (2003) and Mac Low \\& Klessen (2004). In this proceedings paper we call your attention to the dynamical complexity arising from the interplay between supersonic turbulence and self-gravity, and introduce the concept of gravoturbulent fragmentation. We argue that in typical star forming clouds turbulence generates the density structure in the first place and then gravity takes over in the densest and most massive regions to build up the star cluster. In Section 2 we focus on spatial distribution and timescale of star formation, then in Section 3, we discuss a specific example of a star forming filament similar to those observed in Taurus, and in Section 4 we speculate about the mass spectra of clumps and stars in the context of the gravoturbulent fragmentation model. Finally, in Section 5 we demonstrate that the equation of state (EOS) of the interstellar gas plays a pivotal role in gravoturbulent fragmentation. The EOS determines whether molecular cloud regions build up clusters of low to intermediate-mass stars, or form isolated high-mass objects. ", "conclusions": "" }, "0403/astro-ph0403143_arXiv.txt": { "abstract": "{ We present 2\\farcs4 resolution, high sensitivity radio continuum observations of the nearby spiral galaxy NGC\\,4258 in total intensity and linear polarization obtained with the Very Large Array at $\\lambda$3.6~cm (8.44~GHz). \\\\ The radio emission along the northern jet and the center of the galaxy is polarized and allows investigation of the magnetic field. Assuming energy-equipartition between the magnetic field and the relativistic particles and distinguishing between (1) a relativistic electron-proton jet and (2) a relativistic electron-positron jet, we obtain average magnetic field strengths of about (1) 310~$\\mu$G and (2) 90~$\\mu$G. The rotation measure is determined to range from $-$400 to $-$800~rad/m$^{2}$ in the northern jet. Correcting the observed E-vectors of polarized intensity for Faraday rotation, the magnetic field along the jet turns out to be orientated mainly along the jet axis. An observed tilt with respect to the jet axis may indicate also a toroidal magnetic field component or a slightly helical magnetic field around the northern jet. ", "introduction": "The nearby galaxy NGC\\,4258 (M\\,106) is a bright SAB(s)bc spiral (de Vaucouleurs et al.\\ \\cite{vauc}) at a distance of 7.2~Mpc (Herrnstein et al.\\ \\cite{herr99}). It seems to possess a small bright nucleus with a highly excited emission line spectrum (Burbidge et al.\\ \\cite{bur}) and has been classified as a weakly active Seyfert~2-type galaxy. Most striking are the two so-called `anomalous arms', which are not visible in the optical and were first detected in H$\\alpha$ by Court\\`{e}s \\& Cruvellier (\\cite{cour}) in the inner region of the galaxy. Van der Kruit et al. (\\cite{kruit}) detected these anomalous arms in the radio range where they extend out to the optical periphery. Spectral index studies indicate that their radio emission is of non-thermal origin (de Bruyn\\ \\cite{bruyn}; van Albada \\cite{alb}). The anomalous arms of NGC\\,4258 have been extensively discussed in terms of ejection of matter from the nucleus. The detection of a water-maser (Claussen et al.\\ \\cite{claus}; Henkel et al.\\ \\cite{henk}) and an accretion disk around a supermassive central object (Miyoshi et al.\\ \\cite{miyo}; Herrnstein et al.\\ \\cite{herr}), and the fact that inner anomalous arms are orientated parallel to the rotation axis of the accretion disk and can be traced even on subparsec scale (Herrnstein et al.\\ \\cite{her}) indicate that the anomalous arms indeed are jets. Whereas in the inner regions the anomalous arms clearly reveal their jet character, many of the features at a greater distance from the nucleus (e.g. their bifurcation) remain unexplained. The three-dimensional geometry of the galaxy and its jets has been discussed for a long time (cf. e.g. van Albada \\& van der Hulst\\ \\cite{albhulst}; Hummel et al.\\ \\cite{hum}). The detection of an accretion disk revealed directly for the first time that the central part of the galaxy has a significant tilt with respect to the galactic disk: the accretion disk itself has an inclination angle of $83\\degr$ and a position angle (p.a.) of $86\\degr$ (Miyoshi et al.\\ \\cite{miyo}), i.e. it is oriented nearly east-west. The p.a. of the galactic disk, however, is $150\\degr$, which is nearly north-south, and its inclination is $72\\degr$ (van Albada \\cite{albada}). Thus, the plane of the galactic disk and the plane of the accretion disk have a significant angle to each other. As the jets emerge perpendicular to the accretion disk, they have to pass the galactic disk at a rather small angle. Due to the projection of the whole system with respect to the Earth, the inner jet direction is almost parallel to the major axis of the galactic disk. To further investigate the jet geometry and especially the magnetic field along the jets we obtained data with the Very Large Array (VLA)\\footnote{The VLA is a facility of the National Radio Astronomy Observatory. The NRAO is operated by Associated Universities, Inc., under contract with the National Science Foundation.} at $\\lambda$3.6~cm and compared these with already published but reprocessed VLA data by Hummel et al. (\\cite{hum}) at $\\lambda$6.2~cm (4.86~GHz) and $\\lambda$20~cm (1.49~GHz) and with observations at $\\lambda$2.8~cm (10.55~GHz) made with the Effelsberg 100-m telescope. The observations and data reduction procedures are described in Sect.~2. In Sect.~3 we present the results and examine the measurements of the linearly polarized emission in terms of Faraday rotation, magnetic field strength and direction. The discussion of the results and the summary follow in Sect.~4 and Sect.~5, respectively. ", "conclusions": "We present interferometer data obtained with the VLA in its C-configuration at $\\lambda$3.6~cm (8.4399~GHz) in total power and linear polarization. For comparison and to obtain quantities like rotation measure and depolarization we also reprocessed VLA data at $\\lambda$6.2~cm (4.8851~GHz) and $\\lambda$20~cm (1.4899~GHz) that were previously published by Hummel et al. (\\cite{hum}) and observations at $\\lambda$2.8~cm (10.55~GHz) made with the Effelsberg 100-m telescope. The high resolution maps ($2\\farcs4$ HPBW) at $\\lambda$3.6~cm as obtained with the C-configuration of the VLA are able for the first time to resolve the jets in the central region of the galaxy. Detailed inspection shows that they emerge from the galactic center along the projected spin axis of the accretion disk as determined by Miyoshi et al. (\\cite{miyo}). At a distance $24\\arcsec$ away from the center they change direction towards the previously seen northwest to southeast and bifurcate at $62\\arcsec$ from the nucleus in the northern jet and at $85\\arcsec$ in the southern jet. The multiple splitting of the northern jet is clearly visible at this resolution. The polarized emission was detected exclusively along the jets and allowed the calculation of the magnetic field strength in the inner region of the northern jet. Energy equipartition considerations lead to magnetic field strengths of $310\\pm 15~\\mu$G assuming a relativistic electron-proton jet and $90\\pm 5~\\mu$G assuming an electron-positron jet. The rotation measure could be determined between $\\lambda$3.6~cm and $\\lambda$6.2~cm, at a linear resolution of $14\\arcsec$ HPBW. It is derived to be $-$400 to $-$800~rad/m$^2$ in the northern jet. Correcting the observed E-vectors of polarized emission for Faraday rotation, the magnetic field is mainly along the jet axis in the central region and tends to become somewhat tilted with respect to the jet direction in the outer part of the northern jet. This may be consistent with a slightly helical magnetic field around the northern jet or may indicate a superposition of a longitudinal magnetic field near the jet axis and a toroidal magnetic field away from the axis." }, "0403/astro-ph0403625_arXiv.txt": { "abstract": "{\\it RXTE} ASM count rates from the X-ray pulsar Her X-1 began falling consistently during the late months of 2003. The source is undergoing another state transition similar to the anomalous low state of 1999. This new event has triggered observations from both space- and ground-based observatories. In order to aid data interpretation and telescope scheduling, and to facilitate the phase-connection of cycles before and after the state transition, we have re-calculated the precession ephemeris using cycles over the last 3.5 years. We report that the source has displayed a different precession period since the last anomalous event. Additional archival data from \\cgro\\ suggests that each low state is accompanied by a change in precession period and that the subsequent period is correlated with accretion flux. Consequently our analysis reveals long-term accretion disk behaviour which is predicted by theoretical models of radiation-driven warping. ", "introduction": "\\label{sec:introduction} Her X-1 is an eclipsing X-ray pulsar that displays variability on spin (1.24s), orbital (1.7d) and super-orbital (35d) timescales \\citep{tan72}. The companion star is of early-F or late-A type and fills its Roche lobe, resulting in accretion mainly by Roche lobe overflow through the inner Lagrangian ($L_1$) point \\citep{lea98}. The 35d period is revealed in X-rays as an absorbing column of variable depth passing in front of the accretion source \\citep{sco00}, In the UV and optical, an intrinsic flux variation is observed on the same period \\citep{ger76,lea99}. The long-standing model is of a warped accretion disk, viewed close to edge-on ($i \\simeq 90$ deg) precessing in a retrograde direction around the neutron star, \\citep{kat73,rob74}. The disk atmosphere provides the column and casts an X-ray shadow over the inner face of the companion star, causing optical variability on 35d timescales. \\begin{figure*} \\begin{picture}(100,0)(10,20) \\put(0,0){\\special{psfile=\"f1_color.eps\" hoffset=20 voffset=-550 angle=0 vscale=80 hscale=80}} \\noindent \\end{picture} \\vspace{190mm} \\figcaption{Top panel: \\her\\ \\cgro\\ BATSE (20--50 \\kev) one-day average count rates from Earth occultations between 1991--1999. Middle panel: \\xte\\ ASM (1--12 \\kev) one-day average count rates from \\her\\ from 1996--2004. Bottom panel: ASM Hardness ratio averaged over each 35 day cycle. Band A = 1.3--3 \\kev, band B = 3--5\\kev\\ and band C = 5--12\\kev. Horizontal lines represent count rate thresholds adopted in the $0-C$ analysis. Dwells over the two thresholds are plotted on the $O-C$ diagram of Fig. \\ref{fig:o-c}\\label{fig:asm}} \\end{figure*} In X-rays, the 35d cycle is characterized by a rapid turn-on, a ``main-on'' state which varies in intensity greatly from cycle to cycle, and then decays back to an ``off'' state at $\\sim 1$\\% of the peak main-on flux. This is followed by a rise to a weaker ``short-on'' state and then another low \\citep{tan72,gor82}. Superimposed on this cycle are eclipses of the neutron star by the companion once per orbit and dips that occur almost on the beat period between orbit and precession \\citep{gia73}, which are believed to be due to localized structure in the disk, generated by the impact of the ballistic gas stream from the $L_1$ point \\citep{cro80,sch96}. The turn-on occurs at only two orbital phases, $\\phio \\simeq 0.2$ or $\\phio \\simeq 0.7$, apparently at random. Consequently it has often been claimed that precession is not strictly periodic (Ogelman 1987), and precession cycles generally have durations of either 20, 20.5 or 21 orbital cycles, leading some authors to suggest that the precession and orbital cycles are related physically \\citep{sco99}. To date, it has not been possible to determine the engine behind disk precession observationally. From a theoretical perspective, the companion star can force a tilted disk to precess in the outer regions closest to the donor star \\citep{lar98}. However tidal forces do not account for the mechanism that tilts the disk. Mechanisms that can simultaneously warp structure and drive precession are wind \\citep{sch94} and radiation \\citep{pri96,wij99} pressure, pushing down on the disk plane. The dominant engine behind this pressure is accretion flux from the neutron star which irradiates the disk atmosphere externally. While the simplifying assumptions used in these theoretical calculations prevent a detailed comparison with observation, the success of the models in producing warped disks with the observed precession rate indicates that some combination of tides, wind and radiation pressure are capable of driving the \\her\\ clock. Her X-1 has occasionally displayed consecutive main-on states of significantly lower flux than average, e.g. 1993 \\citep{vrt94}. Also, between 1999--2000, the source missed a large number of consecutive cycles altogether \\citep{par99,cob00,vrt01,sti01}. See also \\citet{par85} for the description of a 1983 event prior to the epoch of all-sky monitoring surveys. During these times, the optical flux does not dip appreciably \\citep{vrt01}. In addition to the decrease in main-on amplitude, the neutron star experiences an abrupt transition from the nominal state of pulse period spin-up to episodes of spin-down \\citep{par99}. Although these events are not homogeneous in terms of either duration or magnitude, they have been coined collectively the `Anomalous Low States (ALS)'. Quite clearly the ALS are significant disk events; they point to the accretion disk changing state for a limited period of time before returning seemingly to the same preferred period and warp shape, albeit with an offset of several orbital cycles in the turn-on clock, according to \\citet{oos01} and \\citet{man03}. In this paper, we show this picture to be incorrect. {\\it Rossi X-ray Timing Explorer (RXTE)} All-Sky Monitor (ASM) count rates from Her X-1 in the main-on state began falling consistently during the late months of 2003. \\her\\ appears to be undergoing another state transition similar to the ALS of 1999--2000. Since a glitch in the precession clock was reported after the last ALS \\citep{oos01}, it is now prudent to recalculate the precession ephemeris using the last 3.5 years of data. Rather than a glitch, we report that the source has displayed a different precession period, \\pprec, since the last ALS. Additionally, archival data from the BATSE experiment that was onboard the {\\it Compton Gamma Ray Observatory (CGRO)} suggests that each ALS is accompanied by a change in \\pprec. Contrary to previous reports \\citep{oos01}, these results indicate that the accretion disk does not return to the same state after each ALS. ", "conclusions": "\\label{sec:conclusions} We have redefined the accretion disk precession ephemeris for the X-ray pulsar \\her\\ between the end of the 1999--2000 ALS and the start of the 2003--2004 ALS event. Relative to the most recently published work on cycle timing \\citep{man03}, the new ephemeris provides a correction to the turn-on time of the precession cycle of $\\sim 14$d at the time of writing, critical to research teams and observatory schedulers during the current ALS. A period analysis of the \\cgro\\ BATSE and \\xte\\ ASM time series since 1991 reveals four epochs of stable precession period. The precession epochs are each separated by an ALS and the change of period appears to be instantaneous within the resolution of the $O-C$ diagram, i.e. a few precession cycles. Testable predictions are that \\her\\ will return from the 2003--2004 ALS with a new precession period and epoch-averaged peak main-on flux." }, "0403/astro-ph0403413_arXiv.txt": { "abstract": "We present a systematic spectral analysis of six ultraluminous X-ray sources (NGC1313 X-1/X-2, IC342 X-1, HoIX X-1, NGC5408 X-1 and NGC3628 X-1) observed with {\\sl XMM-Newton} Observatory. These extra-nuclear X-ray sources in nearby late-type galaxies have been considered as intermediate-mass black hole candidates. We have performed Monte-Carlo simulations of Comptonized multi-color black-body accretion disks. This unified and self-consistent spectral model assumes a spherically symmetric, thermal corona around each disk and accounts for the radiation transfer in the Comptonization. We find that the model provides satisfactory fits to the {\\sl XMM-Newton} spectra of the sources. The characteristic temperatures of the accretion disks ($T_{in}$), for example, are in the range of $\\sim 0.05-0.3$ keV, consistent with the intermediate-mass black hole interpretation. We find that the black hole mass is typically about a few times $10^3~\\rm{M_\\odot}$ and has an accretion rate $\\sim 10^{-6} - 10^{-5} ~\\rm{M_\\odot~yr^{-1}}$. For the spectra considered here, we find that the commonly used multi-color black-body accretion disk model with an additive power law component, though not physical, provides a good mathematical approximation to the Monte-Carlo simulated model. However, the latter model provides additional constraints on the properties of the accretion systems, such as the disk inclination angles and corona optical depths. ", "introduction": "Probably the most exciting recent development in the field of black hole (BH) study is the discovery of numerous candidates for intermediate-mass BHs (IMBHs) with masses in the range of $\\sim 10^2-10^5 M_\\odot$ (see Miller \\& Colbert 2003 for a recent review). The presence of such BHs was first proposed to explain ultraluminous X-ray sources (ULXs), defined as point-like extra-nuclear X-ray sources observed in nearby galaxies and with inferred isotropic X-ray luminosities in excess of $10^{39} {\\rm~erg~s^{-1}}$, about an order of magnitude greater than the Eddington limit of a solar mass object (e.g., Fabbiano 1989; Colbert \\& Mushotzky 1999; Miller et al. 2003a,b; Strohmayer \\& Mushotzky 2003). While unambiguous detections of individual IMBHs currently do not exist, there are observational hints from studies of microlensing events, globular clusters, and centers of nearby galaxies (van der Marel 2003 and references therein). Although ULXs actually represent a heterogeneous population, a majority of them are likely to be accreting BHs. The controversy is centered on the X-ray emission mechanisms and on the masses of the BHs (e.g., Makishima et al. 2000; King et al. 2001; Begelman 2002). The IMBH interpretation, though probably the most straightforward and exciting, has serious difficulties (e.g., King et al. 2001; Kubota et al. 2002). In addition to their high X-ray luminosities, many ULXs show convex-shaped spectra, especially in the energy band $\\lesssim $ a few keV. Such spectra are characteristic of the ``soft state'' of accreting BH binaries and are often approximated by black-body-like models such as the multi-color disk model (MCD; {\\em diskbb} in the {\\em XSPEC} spectral analysis software package (Arnaud 1996; e.g., Makishima et al. 1986). In the MCD model, each annulus of the axis-symmetric optically-thick accretion is assumed to radiate as a blackbody with a radius-dependent temperature. The characteristic temperature $T_{in}$ of the innermost portion of the disk is $\\propto (\\dot{M}/M)^{1/4}$, where $M$ and $\\dot{M}$ are the BH mass and the accretion mass rate. However, $T_{in}$ inferred from the model fit is almost always too high for the required high mass, assuming the Eddington limit on $\\dot{M}$. Equivalently, the inner disk radius $R_{in}$ is much smaller than the last stable orbit for a non-spin BH. Even more disturbing is that the inferred value of $R_{in}$ is sometimes found to be time-variable, in contrast to the soft-state of confirmed BH X-ray binaries, where $R_{in}$ is approximately constant for each source. Furthermore, it has been shown recently that many X-ray spectra of ULXs cannot be satisfactorily described by a single MCD model, especially in observations with good counting statistics and with a broad energy coverage (e.g., Miller et al. 2003a,b). The usual practice is then to fit such a spectrum with an additive combination of an MCD and a power law (PL; e.g, Miller 2003a,b; Cropper \\etal 2003). The requirement of this latter component, which often becomes important at energies $\\gtrsim$ a few keV and may extend up to 200 keV as indicated by stellar mass BH systems, suggests that a high temperature electron cloud (corona) exists around an accretion disk, producing inverse-Compton scattering of the disk photons (e.g., Kubota et al. 2002; Page et al. 2003). A spectral fit with this additive, phenomenological model combination typically leads to an acceptable fit and a much lower (apparently more reasonable) disk temperature. Nevertheless, the model is over-simplified in the following two aspects: First, the extension of the power-law component straight to the low-energy limit of the spectrum is nonphysical. Because the power-law component is assumed to mimic the effect of the inverse Compton scattering, a low-energy cutoff as in the MCD component must be present in the Comptonized component (see also Page et al. 2003). Neglecting this low-energy cutoff could mis-characterize the spectral shape of the MCD component and could lead to an artificially high absorption in spectral fitting. Second, the additive combination of the MCD and the PL components does not account for the radiative transfer process or the removal of photons from the MCD component to the Comptonized component. Furthermore, this process depends on both disk photon and corona electron energies. Therefore, one may not, in general, directly take the MCD normalization derived from the spectral model fitting to infer the inner disk radius or the BH mass, as realized by some authors (e.g., Kubota, Makishima \\& Ebisawa 2001). Otherwise, the inferred (but probably nonphysical) inner disk radius, for example, may appear to vary significantly when an accretion system changes from its hard state to its soft state or vice versa, as in XTE J2012+381 (Campana et al. 2002). These oversimplifications in the MCD+PL model certainly obscure the physical dependence of the Comptonization on the corona and disk properties, and could also seriously affect the inference of accretion disk parameters. Yao \\etal (2003, hereafter Paper I) have recently developed a Comptonized multi-color disk (CMCD) model. They use Monte-Carlo simulations to directly generate the Comptonized X-ray spectra, removing the above mentioned over-simplifications and avoiding the complications of using the two~(unrelated)-component model and then trying to correct for various radiative transfer effects. While the MCD model is still used to describe the accretion disk emission, the Comptonized radiation is no longer an independent component. This self-consistent treatment thus provides a new tool to constrain the physical properties of the corona and to study its relationship to the accretion disk. But most importantly, the model enables us to recover the same original disk flux in a spectral fit, which is essential to a reliable mass estimate of the putative BH. In the present work, we apply the CMCD model to the analysis of {\\sl XMM-Newton} spectra of six ULXs which have been suggested as IMBH candidates. Our main objectives are to check whether or not the CMCD model provides an adequate spectral description of these sources and to see what potential new insights we may gain from such an application. The sources and data are described in \\S 2, whereas the implementation of the model for this application is discussed in \\S 3, which also includes a summary of various corrections required to infer BH masses from the present model fits. We also test the PL, MCD, and MCD+PL models and compare them with the CMCD model. We present the results of our spectral fits in \\S 4 and discuss the implications and conclusions in \\S 5. ", "conclusions": "The satisfactory fits of the CMCD model to the {\\sl XMM-Newton} spectra of our selected ULXs suggest that they are consistent with the IMBH interpretation. In particular, the model does not have the high $T_{in}$ problem as is faced by the MCD model. The problem is apparently caused by the neglect of Comptonization in the model. Although this neglect is statistically allowed when both the counting statistics and the energy band coverage of an observed spectrum are poor, the fitted spectral parameters are far from being physical. We conclude that the MCD model alone should not be used in the interpretation of ULXs as IMBHs. We confirm that both $T_{in}$ and $N_H$ inferred from the MCD+PL model are reasonably accurate for the sources considered here. This apparent agreement between the CMCD and MCD+PL models suggests that the latter model as a whole is mathematically a good representation of the former model, at least for the IMBH candidates considered here. This is rather surprising when one considers the over-simplifications in the MCD+PL model, as discussed in \\S 1. It appears that the nonphysical extrapolation of the PL to the low energy nearly compensates the failure to include the radiation transfer loss of soft disk photons. However, this does not mean that the MCD+PL model could be used to describe the Comptonized disk emission in general. The various nonphysical effects can cause problems for other sources, especially those with higher $T_{in}$ values ($\\sim$ 1 keV; see the discussion in \\S 3; e.g., LMC~X-1 and LMC~X-3; Paper II). In comparison, the CMCD model provides more reliable measurements of the disk parameters as well as unique constraints on the physical properties of the coronae. In the following, we briefly discuss both the function of these new parameters and the physical reason for their different degrees of constraints: \\begin{itemize} \\item The opacity $\\tau$ is relatively well constrained, which is the key parameter that determines the total number of Comptonized photons (e.g., Fig.~\\ref{fig:compare}). For example, IC 342 X-1 with the largest best-fit $\\tau$ value appears to have the disk emission nearly completely Comptonized, explaining why the spectrum of the source can be characterized by a PL alone. Also for this source, because the saturated Comptonization dominates over the thermal emission, the constraints on $T_{in}$, $R_{in}$, and eventually on BH mass are very weak. \\item The corona electron temperature $T_c$ is chiefly responsible for the overall energy extent of the Comptonized spectral component. Because $T_c$ is $\\gtrsim 30$ keV for all the sources, the high-energy turning-off of the component is well beyond the \\xmm\\ band limit. Therefore, the data do not constrain the upper limit of $T_c$. The lower limit is determined because a minimum electron energy is needed to up-scatter soft disk photons to the high energies covered by the spectra. \\item Whereas the nearly isotropic Comptonized flux is barely affected by the disk inclination angle $\\theta$, the observed strength of the soft disk component is proportional to cos($\\theta$). In a spectral fit, however, this difference in the disk inclination dependence may be partially compensated by a change in the $\\tau$ value. But if $\\theta$ is large (for a nearly edge-on disk), its geometric effect cannot be canceled by adjusting other parameters, which would also effectively alter the spectral shapes of both the disk and Comptonized components in a spectral fit (e.g., Fig.~\\ref{fig:compare}). Consequently, we may constrain the upper limit to $\\theta$. This constraint, though not very tight, is important for the estimation of the BH masses (\\S 3). \\item $R_c$ determines the effective corona radius, within which the disk emission is most affected by the Comptonization. Photons from larger radii have relatively little chance to be scattered and may contribute to the un-Comptonized disk component even if $\\tau$ is large ($\\geq 1$). But the amount of soft X-ray radiation from the disk also decreases with the increasing radius. The combination of these two effects may thus place a constraint on $R_c$. \\end{itemize} Apparently, these parameters are correlated in a spectral fit. This, together with the limited counting statistics and bandwidth of the data, explains why the parameters are not tightly constrained. Nevertheless, the results presented above demonstrate the potential of the CMCD model to shed new insights into the physical properties of the accretion disk coronae, in addition to a more reliable mass estimate of the putative BHs. Table~\\ref{results:flux} includes our estimated BH masses, assuming no spin and the best-fit $\\theta$ values. The typical BH mass is in the range of $\\sim 10^3-10^4$ M$_\\odot$, although the upper limit for IC~342~X-1 is slightly higher. If a BH spins rapidly, the inferred BH mass could be several times higher than the value quoted in the table (depending on $\\theta$; Fig. 2). Assuming that the bolometric luminosity (estimated in the 0.05--100 keV range) $L_{bol} = 0.1 \\dot{M}c^2$, we further estimate the accretion rate $\\dot{M}$ for each source (Table 5), which is in the range of 1--10 $\\times 10^{-6}$ M${\\rm_{\\odot}yr^{-1}}$. The present work represents, at most, an incremental step in developing a fully self-consistent model for accreting BH systems. The CMCD model used here deals only with the Comptonization by static disk coronae. To study the dynamics, one needs to understand the formation and evolution of the coronae as well as the physics of the accretion disks. We also have not considered other proposed scenarios that may explain some of the ULXs; e.g., the anisotropic emission of the radiation \\citep{king01}, the relativistic motion of the X-ray-emitting plasmas \\citep{fab01}, and the possible super-Eddington emission (Begelman 2002). Spectral models for such scenarios, yet to be developed, need to account for the apparent presence of the soft thermal component, in addition to the power law, for these sources except in the case of IC 342~X-1. The bottom line here is that the \\xmm\\ spectra are consistent with the IMBH interpretation of the sources." }, "0403/astro-ph0403139_arXiv.txt": { "abstract": "We present a deep JHK$_s$-band imaging survey of the W3 Main star forming region, using the near-infrared camera, SIRIUS (Simultaneous three-color InfraRed Imager for Unbiased Surveys), mounted on the University of Hawaii 2.2 m telescope. The near-infrared survey covers an area of $\\sim$ 24 arcmin$^2$ with 10 $\\sigma$ limiting magnitudes of $\\sim$ 19.0, 18.1, and 17.3 in J, H, and K$_s$-band, respectively. We construct JHK color-color and J/J-H and K/H-K color-magnitude diagrams to identify young stellar objects and estimate their masses. Based on these color-color and color-magnitude diagrams, a rich population of YSOs is identified which is associated with the W3 Main region. A large number of previously unreported red sources (H-K $>$ 2) have also been detected around W3 Main. We argue that these red stars are most probably pre-main sequence stars with intrinsic color excesses. We find that the slope of the K$_s$-band luminosity function of W3 Main is lower than the typical values reported for the young embedded clusters. The derived slope of the KLF is the same as that found by Megeath et al. (1996), from which analysis by Megeath et al. indicates that the W3 Main region has an age in the range of 0.3--1 Myr. Based on the comparison between models of pre-main sequence stars with the observed color-magnitude diagram we find that the stellar population in W3 Main is primarily composed of low mass pre-main sequence stars. We also report the detection of isolated young stars with large infrared excesses which are most probably in their earliest evolutionary phases. ", "introduction": "The W3 giant molecular cloud (GMC) complex is located in the Perseus spiral arm at a distance of 1.83$\\pm$0.14 kpc from the Sun (Imai et al. 2000). W3 GMC hosts two massive and active star forming regions, W3 Main in the north, and W3 (OH) in the south. The W3 Main star forming region contains objects such as H II regions, embedded infrared sources (including the extremely luminous cluster of sources W3 IRS 5), OH and water masers (Wynn-Williams et al. 1974, Forster et al. 1977, Gaume \\& Mutel 1987), which are in different stages of evolution. The millimeter continuum observations have shown the existence of two dense clumps of about 2000 M$_{\\odot}$ associated with the luminous infrared sources IRS 4 and IRS 5 in the W3 core (Richardson et al. 1989). VLA observations have resolved the IRS 5 region into a cluster of seven distinct centimeter radio sources (Tieftrunk et al. 1997, hereafter TGC97; Claussen et al. 1994, hereafter CG94). TGC97 proposed that the spatial and kinematic relation of the compact, ultracompact, and hypercompact radio continuum regions toward W3 Main is indicative of sequentially triggered star formation caused by the pressure of the expanding H II regions and the subsequent compression of the molecular gas. Recently, high resolution continuum imaging at 1.3 and 0.7 cm of four hypercompact H II regions in W3 IRS 5, suggested that these sources contain B or O stars (Wilson et al. 2003). From a recent near-infrared (NIR) survey of a $\\sim$ 1\\arcmin.5$\\times$1\\arcmin.5 region towards W3 Main, Megeath et al. (1996) found a dense concentration of stars in the molecular clump surrounding W3 IRS 5. These data showed a large, embedded population of intermediate to low mass stars co-existing with recently formed OB stars. They also argued that the formation of high mass stars is associated with the formation of dense clusters of low mass stars in the W3 Main star forming region. Several X-ray sources were detected in the W3 core by Chandra X-ray Observatory (Hofner et al. 2002). Most of these sources are located at the peak radio positions of the W3 H II regions. Hofner et al. (2002) postulated that the X-ray sources are the young massive stars that are also responsible for the ionization of the compact and ultracompact H II regions in the W3 core. In this paper we present deep J, H, and K$_s$-bands NIR observations of the W3 Main star forming region. In comparison with the previous NIR survey (Megeath et al. 1996), our survey covers a larger area ($\\sim$ 24 arcmin$^2$) surrounding W3 IRS 5, including compact H II regions W3 A, W3 B and W3 D, diffuse H II regions W3 H, W3 J and W3 K, and the cometary ultracompact (UC) H II regions W3 C, W3 E and W3 F. These individual H II regions have been labeled following the scheme introduced by Wynn-Williams (1971) and Harris \\& Wynn-Williams (1976). Our motivation is to look for new young stellar objects (YSOs) associated with the W3 Main region, to determine their evolutionary stages, and to discuss their nature. Tieftrunk et al. (1998) have presented the three 10\\arcmin$\\times$10\\arcmin ~mosaics ($\\sim$ 300 arcmin$^2$) in K' filter of W3 region. This mosaic has similar spatial resolution and depth to the SIRIUS K$_s$ image and covers the region stretching from W3 Main to W3(OH). However, due to the crowded nature of the sources and the large pixel size, the K'-band mosaic was not suitable for photometry of the embedded stellar clusters. In Sects. 2 and 3 we present the details of observations and data reduction procedures, Sect. 4 deals with the results and discussion and we summarize our conclusions in Sect. 5. ", "conclusions": "A deep JHK$_s$-band NIR imaging survey of YSOs associated with the W3 Main star forming region is presented. The survey covers a 4\\arcmin.9$\\times$4\\arcmin.9 area down to a limiting magnitude (10 $\\sigma$) of J = 19.0, H = 18.1, and K$_s$ = 17.3. From the analysis of these images we derive the following conclusions : 1) A cluster of YSOs (Class II and Class I sources derived from their NIR colors) has been detected in the W3 Main core and near the compact, ultracompact, and diffuse H II regions. 2) A large number of red stars (H-K $>$ 2) are detected in the molecular cloud region, most of them clustered around the molecular clumps associated with IRS 5 and IRS 4. Some of them are also associated with the diffuse emission near the dense molecular clumps. We argue that most of the reddest stars are YSOs with circumstellar materials. 3) The KLF of the W3 Main region shows the power-law slope : $\\alpha$ = 0.26$\\pm$0.02, which is lower than the typical values reported for the embedded young clusters. Our finding also confirms the previous results of Megeath et al. (1996) for a smaller region around W3 IRS 5. 4) The observed density of the cluster region around W3 IRS 5 is $\\sim$ 2000 stars pc$^{-3}$ for K $<$ 17.5, which is larger than the typical values ($\\sim$ 1000 stars pc$^{-3}$) reported for other embedded clusters. 5) Using the age of W3 Main in the range of 0.3--1 Myr determined by Megeath et al. (1996), we find that about 80\\% of the YSO candidates have an upper mass limit of 4 M$_{\\odot}$. We estimate that the lowest mass limit of Class II \\& Class I candidates in our observations is 0.1 M$_{\\odot}$. Therefore, the stellar population in W3 Main is primarily composed of low mass PMS stars." }, "0403/astro-ph0403280_arXiv.txt": { "abstract": "After a brief introduction to the necessary of {\\em quark stars} in modelling pulsars, I present a qualitative analysis of the {\\em solidification} of quark matter with low-temperature but high-density. The reason, that a solid neutron star could {\\em not} be possible, is also given. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403555_arXiv.txt": { "abstract": "{ We analyse the evolution of the fractional ionisation in a steady-state protoplanetary disc over $10^6$~yr. We consider a disc model with a vertical temperature gradient and with gas-grain chemistry including surface reactions. The ionisation due to stellar X-rays, stellar and interstellar UV radiation, cosmic rays and radionuclide decay is taken into account. Using our reduction schemes as a tool for the analysis, we isolate small sets of chemical reactions that reproduce the evolution of the ionisation degree at representative disc locations with an accuracy of 50\\%--100\\%. On the basis of fractional ionisation, the disc can be divided into three distinct layers. In the dark dense midplane the ionisation degree is sustained by cosmic rays and radionuclides only and is very low, $\\la 10^{-12}$. This region corresponds to the so-called ``dead zone'' in terms of the angular momentum transport driven by MHD turbulence. The ionisation degree can be accurately reproduced by chemical networks with about 10 species and a similar number of reactions. In the intermediate layer the chemistry of the fractional ionisation is driven mainly by the attenuated stellar X-rays and is far more complicated. For the first time, we argue that surface hydrogenation of long carbon chains can be of crucial importance for the evolution of the ionisation degree in protoplanetary discs. In the intermediate layer reduced networks contain more than a 100 species and hundreds of reactions. Finally, in the unshielded low-density surface layer of the disc the chemical life cycle of the ionisation degree comprises a restricted set of photoionisation-recombination processes. It is sufficient to keep about 20 species and reactions in reduced networks. Furthermore, column densities of key molecules are calculated and compared to the results of other recent studies and observational data. The relevance of our results to the MHD modelling of protoplanetary discs is discussed. ", "introduction": "Nowadays a paradigm of the evolution of a protoplanetary disc is widely accepted in which most of the disc matter is assumed to move steadily toward a protostar due to redistribution of the angular momentum. Ionisation in such objects is an important factor that enables the angular momentum transport to occur via magnetohydrodynamic (MHD) turbulence driven by the magnetorotational instability (MRI; Balbus \\& Hawley~\\cite{MRI}). From this point of view, a disc is conventionally divided into the ``active'' layer, adjacent to the disc surface, and the ``dead'' zone, centred on the midplane. The active part of the disc is irradiated by high energy stellar/interstellar photons. Thus, the fractional ionisation there is relatively high, which implies that the magnetic field is well coupled to the gas. Due to this coupling, the active layer is unstable to the MRI, and the developing turbulence allows the accretion to occur (e.g., Gammie~\\cite{gammie}; Fleming \\& Stone~\\cite{fs2003}). The shielded midplane region is almost neutral, decoupled from the magnetic field, and, thus, quiescent. The location of the boundary between these two regions may prove to be very sensitive to the disc physical properties and chemical composition (e.g., Fromang, Terquem, \\& Balbus~\\cite{from2002}). However, the MHD modelling with non-ideal effects included is a very demanding computational task. This is why in the MHD modelling of protoplanetary discs (and protostellar clouds) a very simple chemical scheme is usually assumed with a few ions and a network that includes only ionisation and recombination reactions, often neglecting the presence of dust grains (Sano et al.~\\cite{sano}; Fromang et al.~\\cite{from2002}; Fleming \\& Stone~\\cite{fs2003}). The medium is believed to be in chemical equilibrium so that the fractional ionisation $x_\\mathrm{e}$ can be expressed as (e.g., Gammie~\\cite{gammie}) \\begin{equation} x_\\mathrm{e}(\\mathrm{eq})=\\sqrt{\\zeta\\over\\beta n_\\mathrm{H}}, \\label{xe} \\end{equation} where $\\zeta$ is the ionisation rate, $\\beta$ is a typical recombination coefficient, and $n_\\mathrm{H}$ is the hydrogen number density. This approach may indeed be valid if only the cosmic ray ionisation is taken into account and only gas-phase chemical processes are considered. However, dust plays an important role in the evolution of the fractional ionisation being an efficient electron donor for recombining ions and a sink for neutrals in cold parts of a disc. Moreover, newly born stars possess a relatively high X-ray flux with photon energies from about 1 to 5 keV (e.g. Igea \\& Glassgold~\\cite{IG99}). These factors give rise to a more complicated chemistry relevant for the ionisation degree. In particular, different ions dominate the fractional ionisation at different times and in different parts of the disc. Therefore, the straightforward application of Eq.~(\\ref{xe}) for evaluating the ionisation degree can be fraught with errors in some parts of a disc. To check the validity of Eq.~(\\ref{xe}), we analyse the detailed ionisation structure of a protoplanetary disc computed with the full UMIST\\,95 chemical network (Millar et al.~\\cite{umist95}) with the surface chemistry included. We adopt a reference disc model to serve as a guide to the range of possible physical conditions that may be encountered in real protoplanetary objects. Within this model, we choose several representative disc locations and investigate in detail which chemical processes control the time-dependent ionisation degree there. Our intention is to show that the oversimplified treatment of the ionisation in a protoplanetary disc as an equilibrated ionisation-recombination cycle can lead to $x_\\mathrm{e}$ values that differ by more than an order of magnitude from values computed with all the available information on the disc chemical evolution. We describe the chemical processes that are responsible for ionisation in a disc in terms of reduced networks that are subsets of the full UMIST\\,95 network, supplied in some cases by a few surface reactions. These networks contain only those species and reactions that are needed to reproduce $x_{\\rm e}$ with up to a factor of 2 uncertainty. The utilised reduction methods are described in Wiebe, Semenov \\& Henning (\\cite{papi}, hereafter Paper~I). The species-based reduction rests upon the Ruffle et al.~(\\cite{Rea02}) technique and consists of choosing species that are important in a particular context, and then selecting from the entire network only those species that are necessary to compute abundances of important species with a reasonable accuracy. In the reaction-based method, analysis starts from reactions that govern the abundance of important species. All reactions in the entire network are assigned weights according to the influence they have on abundances of important species. Then, only those reactions are selected that have weights above a cut-off parameter that is selected on the basis of the requested accuracy. Only those species are included in the reduced network that participate in selected reactions. The organisation of the paper is the following. In Section~\\ref{mod} we describe the disc model and its physical structure as well as updates to the chemical model of Paper~I. In Section~\\ref{icon} we discuss the initial conditions for the disc chemistry. The processes responsible for the fractional ionisation in various parts of the disc are outlined in Section~\\ref{res}. In Section~\\ref{colden} column densities are tabulated and compared to other studies and observational data. Results of the analysis and their relevance to the MHD modelling are discussed in Section~\\ref{diss}. Final conclusions are drawn in Section~\\ref{concl}. ", "conclusions": "\\label{concl} Chemical processes, responsible for the ionisation structure of a protoplanetary disc with a central star, are analysed by means of reduced networks that reproduce the ionisation fraction within a factor of 2. These networks are available from the authors upon request. Because of the wide range of physical conditions met in a typical disc, there is a corresponding diversity in the chemical reactions that control the fractional ionisation in different parts of the disc. Generally, it can be divided into three layers. In the midplane the ionisation is provided by cosmic rays and radioactive elements only. Above the midplane, the intermediate layer is located where the ionisation is dominated by X-rays. In the surface layer UV photons are the main ionising factor. In each of these layers we analyse several representative points and construct reduced chemical networks that are needed to reproduce the fractional ionisation as a function of time during $10^6$ years of evolution within a factor of 2~uncertainty. In the midplane the chemistry, which determines the fractional ionisation, is very simple. Reduced networks typically include no more than 10 species and a similar number of reactions. The same holds true for the surface layer. The intermediate layer is the most complicated region from the chemical point of view. To compute the fractional ionisation with the targeted accuracy, in some regions one has to take into account carbon chains containing up to 6~carbon atoms, which leads to reduced networks with over a hundred species and reactions." }, "0403/astro-ph0403024_arXiv.txt": { "abstract": "The quantum model of the homogeneous and isotropic universe predicts logarithmic-law dependence of coordinate distance to source on redshift $z$ which is in good agreement with type Ia supernovae and radio galaxies observations for the redshift range $z = 0.01 - 1.8$. A comparison with phenomenological models with dark energy in the form of cosmological constant and without dark energy component is made. Fluctuations of the cosmological scale factor about its average value which can arise in the early universe produce accelerating or decelerating expansions of space subdomains containing separate sources with high redshift whereas the universe as a whole expands at a steady rate. ", "introduction": "\\label{Intro} The high-redshift type Ia supernova (SN Ia) observations \\cite{Rie,Per,Ton} can be explained by an accelerating expansion of the present-day universe \\cite{Rie,Per,Tu1,Pee,Liv}. Such conclusion assumes that observed dimming of the SNe Ia is hardly caused by physical phenomena non-related to overall expansion of the universe as a whole (see e.g. Refs. \\cite{Ton,Vish} for discussion and review). Furthermore it is supposed that matter component of energy density in the universe $\\rho_{M}$ varies with the expansion of the universe as $a^{-3}$ (i.e. it has practically vanishing pressure, $p_{M} \\approx 0$), where $a$ is the cosmological scale factor, while mysterious cosmic fluid (so-called dark energy \\cite{Tu2,Ost}) is describes by the equation of state $p_{X} = w_{X} \\rho_{X}$, where $-1 \\leq w_{X} \\leq - \\frac{1}{3}$ \\cite{Tu1,Pee}. In the models with the cosmological constant ($\\Lambda$CDM) one has $w_{X} = - 1$ \\cite{Ton,Tu1,Pee}, while in general case $w_{X}$ may vary with time (models with quintessence) \\cite{Pee,Liv}. Even if regarding baryon component one can assume that it decreases as $a^{-3}$ (pressure of baryons may be neglected due to their relative small amount in the universe), for dark matter (whose nature and properties can be extracted only from its gravitational action on ordinary matter) such a dependence on the scale factor may not hold in the universe taken as a whole (in contrast to local manifestations e.g. in large-scale structure formation, where dependence $a^{-3}$ may survive). Since the contribution from all baryons into the total energy density does not exceed 4 \\% \\cite{Hag}, the evolution of the universe as a whole is determined mainly by the properties of dark matter and dark energy. The models of dark energy \\cite{Tu1,Pee,Tu2} show explicitly unusual behaviour of this component during the expansion of the universe. In the present paper we notice that the quantum model of the homogeneous and isotropic universe filled with primordial matter in the form of the uniform scalar field proposed in Refs. \\cite{K,KK} allows to explain the observed coordinate distances to SNe Ia and radio galaxies (RGs) in wide redshift range. ", "conclusions": "\\label{Con} In addition to the prediction about the steady-speed expansion of the universe as a whole (at the same time the accelerating or decelerating motions of its subdomains remain possible on a cosmological scale as it is shown in Sect. \\ref{Quan}) the quantum model allows an increase of quantity of matter/energy in matter dominated universe according to (\\ref{3}). If the mass $m$ of elementary excitations of the scaler field remains unchanged during the expansion of the universe, then the increase of $M$ can occur due to increase in number $s$ of these excitations. But the increase in $s$ does not mean that a quantity of observed matter in some chosen volume of the universe increases. According to the model proposed in Refs. \\cite{KK2,KK3} the observed ``real'' matter (both luminous and dark) is created as a result of the decay of excitations of the scalar field (under the action of gravitational forces) into baryons, leptons and dark matter. The undecayed part of them forms what can be called a dark energy. Such a decay scheme leads to realistic estimates of the percentage of baryons, dark matter and energy in the universe with $\\langle a \\rangle \\gg 1$ and $M \\gg 1$. Despite the fact that the quantity of matter/energy can increase, the mean total energy density decreases and during the expansion of the universe mainly the number of elementary excitations of the scalar field increases. Their decay probability is very small, so that basically only the dark energy is created. These circumstances can explain the absence of observed events of creation of a new baryonic matter on a cosmologically significant scale. The proposed approach to the explanation of observed dimming of some SNe Ia may provoke objections in connection with the problem of large-scale structure formation in the universe, since the energy density $\\langle \\rho \\rangle$ in the form (\\ref{4}) cannot ensure an existence of a growing mode of the density contrast $\\delta \\langle \\rho \\rangle/\\langle \\rho \\rangle$ (see e.g. Refs. \\cite{Ol,We,Pee2}). As we have already mentioned above in Sect. \\ref{Coor} the density $\\langle \\rho \\rangle$ (\\ref{4}) describes only homogenized properties of the universe as a whole. It cannot be used in calculations of fluctuations of energy density about the mean value $\\langle \\rho \\rangle$. Under the study of large-scale structure formation one should proceed from the more general expression for the energy density (\\ref{2}). Defining concretely the contents of matter/energy $M$, as for instance in the model of creation of matter mentioned above, one can make calculations of density contrast as a function of redshift. The problem of large-scale structure formation is one of the main problems of cosmology (see e.g. Refs. \\cite{Ol,Dol}). It goes beyond the tasks of this paper and requires a special investigation. The ways of its solution in the quantum model are roughly outlined in Ref. \\cite{KK}." }, "0403/astro-ph0403212_arXiv.txt": { "abstract": "Dense high-precision photometry of microlensed stars during a fold-caustic passage can be used to reveal their brightness profiles from which the temperature of the stellar atmosphere as function of fractional radius can be derived. While the capabilities of current microlensing follow-up campaigns such as PLANET allowed for several precise measurements of linear limb-darkening coefficients, all attempts to reveal a second limb-darkening coefficient from such events have failed. It is shown that the residual signal of a second coefficient characterizing square-root limb darkening is $\\sim\\,$25 times smaller which prevents a proper determination except for unlikely cases of very high caustic-peak-to-outside magnification ratios with no adequate event being observed so far or for source stars passing over a cusp singularity. Although the presence of limb darkening can be well established from the data, a reliable measurement of the index of an underlying power-law cannot be obtained. ", "introduction": "Since \\citet{SW1987} have proposed to use caustics of gravitational lenses for measuring the brightness profile of closely-aligned background sources, dense high-precision photometry on microlensing events involving fold-caustic passages has provided measurements of linear limb-darkening coefficients for several stars \\citep{PLANET:M41, PLANET:O23, joint, PLANET:EB5}. For stellar brightness profiles that involve linear, square, or square-root terms in $\\cos \\vartheta$, where $\\vartheta$ is the emergent angle, properties of light curves during fold-caustic passages have been studied by \\citet{Rhie:LD}, while prospects and strategies for determining the linear limb-darkening coefficient have recently been discussed in detail by \\citet{Do:FoldLD}, showing the ability of obtaining precise measurements even with moderate use of the current capabilities. This letter will however show that the measurement of additional coefficients, e.g.\\ corresponding to a square-root law term, is not possible for typical events that involve a fold-caustic passage. Attempts of such measurements \\citep{PLANET:EB5} have resulted in the meaningful measurement of a single linear combination of the involved limb-darkening coefficients only, while a similar result has been obtained for an event where the source passes over a cusp \\citep{Abe}, where however the exhibited differential magnification is modest due to the source size being relatively large compared to the cusp. In contrast, the cusp-passage event discussed by \\citet{PLANET:M28} provides a better size ratio between source and cusp and to date presents the only measurement of both linear and square-root limb-darkening coefficients by microlensing. ", "conclusions": "If the stellar brightness profile $\\xi(\\rho)$ is modelled by a linear superposition of three base profile functions, as for the discussed examples of constant, linear, and square-root terms in $\\cos \\vartheta$, the normalization forces one of these profiles to mediate between the other two, so that this base profile can be approximated by a superposition of the others. With several fractional radii $\\rho$ being probed by the fold caustic for any passage phase $\\eta$ according to ${\\bmath{\\mathcal T}} (\\eta,\\rho)$, the residuals of this approximation are smaller for the caustic profile function $G_\\rmn{f}^\\star(\\eta; \\xi)$ than for the stellar brightness profile $\\xi(\\rho)$, which limits the power of photometric microlensing observations during fold-caustic passages for revealing the stellar brightness profile $\\xi(\\rho)$. Compared to the determination of a linear limb-darkening coefficient, the residual signals of an additional square-root limb-darkening coefficient are $\\sim\\,$25 times smaller, making its measurement with current microlensing follow-up campaigns such as PLANET impossible, unless the caustic-peak-to-exit magnification ratio becomes $F_\\rmn{peak}/F_\\rmn{f}^\\star \\ga 40$ for a (typical) photometric accuracy $\\sigma_\\rmn{f}^\\star/F_\\rmn{f}^\\star = 1.5\\,$\\% at the caustic exit where none of such events have been observed so far, or the source passes over a cusp singularity rather than a fold. If more than one limb-darkening base profile is assumed to contribute, only a single characteristic coefficient can be accurately measured which corresponds to a specific superposition of the corresponding base profiles as found for worked examples \\citep{PLANET:EB5,Abe}." }, "0403/astro-ph0403448_arXiv.txt": { "abstract": "We report here the first evidence for planetesimal infall onto the very young Herbig Be star LkH$_\\alpha$234. These results are based on observations acquired over 31 days using spectroscopy of the sodium D lines, the He I 5876\\AA, and hydrogen H$_\\alpha$ lines. We find Redshifted Absorption Components (RAC) with velocities up to 200 km/s and very mild Blueshifted Absorption Components (BEC) up to 100 km/s in the Na I lines. No correlation is observed between the appearance of the Na I RAC \\& BEC and the H$_\\alpha$ and He I line variability, which suggests that these (Na I RAC \\& BEC) are formed in a process unrelated to the circumstellar gas accretion. We interpret the Na I RAC as evidence for an infalling evaporating body, greater than 100 km in diameter, which is able to survive at distances between 2.0 to 0.1 AU from the star. The dramatic appearance of the sodium RAC and mild BEC is readily explained by the dynamics of this infalling body making LkH$_\\alpha$234 the youngest (age $\\sim$ 0.1 Myr) system with evidence for the presence of planetesimals. ", "introduction": "LkH$_\\alpha$234 (m$_v$=11.9mag) is a star embedded in a nebula associated with the NGC 7129 star forming region and is at a distance of about 1250 parsecs. This is one of the youngest known Herbig Be objects and has an age $\\sim$ 0.1 million years and spectral type of B5 (Fuente et al. 2001). Herbig Ae/Be stars are intermediate mass (2 to 10M$_\\odot$) pre-main-sequence stars (Herbig 1960). A number of Herbig Ae stars have been found to possess circumstellar disks and show both photometric and spectroscopic variability (Herbst et al. 1999). LkH$_\\alpha$234 though much hotter and more massive also exhibits these properties (Fuente et al. 2001, Herbst et al. 1999, Polomski et al. 2002). The star shows photometric variability of 1.2 magnitude in the V band and is known posses a circumstellar disk with the disk mass to the stellar mass ratio (M$_D$/M$_*$) $\\leq$ 0.02. A detailed description of LkH$_\\alpha$234 and its circumstellar region can be found in Fuente et al. (2001, and references therein). Spectroscopic variations in metallic lines (like Na I, Ca II, etc.) in young A type stars like $\\beta$ Pictoris (20 to 100 million years) are often interpreted as the signature of star grazing comets or falling evaporating bodies with sizes of tens of kilometers (Grinin et al. 1996, Smith \\& Terrile 1984, Thebault \\& Beust 2001). The only Herbig Be star where grazing cometary transient events have been inferred so far is HD 100546 (Viera et al. 1999). However, HD100546 is much older ($\\geq$10 Myrs) and cooler (spectral type $\\sim$B9) compared to LkH$_\\alpha$234. High resolution spectroscopic monitoring of very young (0.1 to 1 Myrs) Herbig Be stars are scanty and uncharted. Here we report dramatic variations in the spectroscopic lines of Na I 5890\\AA (D2), and 5896\\AA (D1), He I 5876\\AA, and H$_\\alpha$ of LkH$_\\alpha$234 over a period of 31 days during Oct.-Nov. 2003. The observations consists of high resolution spectra (R=30,000) from the Hobby-Eberly Telescope (HET) using the High Resolution Spectrograph (HRS). Section 2 describes observations and in section 3 we discuss the results and section 4 gives the conclusion. ", "conclusions": "We have presented here high resolution spectroscopic monitoring of a Herbig Be star LkH$_\\alpha$234 in the Na I (D2\\&D1), He I (5876\\AA) and H$_\\alpha$ lines. We found no correlations between the Na I and the He I and H$_\\alpha$ line variability in the five data sets over a period of 31 days (7th Oct. 2003 to 8th Nov. 2003) though we do find a correlation between He I and H$_\\alpha$. Thus the origin of variation in the Na I line (RAC up to 200 km/s and BEC up to 100 km/s seen only on 13th Oct. 2003) is different from those of H$_\\alpha$ and He I. While the H$_\\alpha$ and He I line variations are due to episodic gaseous accretion, the Na I RAC indicate a dramatic transient event of solid body infall on to the star and the BEC as the blown away parts of the solid body not projected against the stellar surface. Considering the Keplerian dynamics and the harsh environment of a B5 star, we estimate that a solid body of size $\\geq$100 km broke up and disintegrated at a distance between 0.1 to 2.0 AU from the star. This makes LkH$_\\alpha$234 the youngest system ($\\sim$0.1 Myrs) with evidence for protoplanetary bodies of asteroidal size. We plan to pursue further spectroscopic monitoring of the star with the HET to determine the frequency of such events and further understand the complicated dynamics of this very young circumstellar environment." }, "0403/physics0403019_arXiv.txt": { "abstract": "In MicroPattern Gas Detectors (MPGD) when the pixel size is below 100 $\\mu$m and the number of pixels is large (above 1000) it is virtually impossible to use the conventional PCB read-out approach to bring the signal charge from the individual pixel to the external electronics chain. For this reason a custom CMOS array of 2101 active pixels with 80 $\\mu$m pitch, directly used as the charge collecting anode of a GEM amplifying structure, has been developed and built. Each charge collecting pad, hexagonally shaped, realized using the top metal layer of a deep submicron VLSI technology is individually connected to a full electronics chain (pre-amplifier, shaping-amplifier, sample \\& hold, multiplexer) which is built immediately below it by using the remaining five active layers. The GEM and the drift electrode window are assembled directly over the chip so the ASIC itself becomes the pixelized anode of a MicroPattern Gas Detector. With this approach, for the first time, gas detectors have reached the level of integration and resolution typical of solid state pixel detectors. Results from the first tests of this new read-out concept are presented. An Astronomical X-Ray Polarimetry application is also discussed. ", "introduction": "The most interesting feature of the Gas Electron Multiplier (GEM) is the possibility of full decoupling of the charge amplification structure from the read-out structure. In this way both can be independently optimized. Indeed, by organizing the read-out plane in a multi-pixel pattern it is possible to get a true 2D imaging capability. At the same time a high granularity of the read-out plane would also allow to preserve the intrinsic resolving power of the device and its high rate capability that otherwise would be unavoidably lost by using a conventional projective read-out approach. However, when the pixel size is small (below 100 $\\mu$m) and the number of pixels is large (above 1000) it is virtually impossible to bring the signal charge from the individual pixel to a chain of external read-out electronics even by using an advanced, fine-line, multi-layer, PCB technology. The fan-out which connects the segmented anodes collecting the charge to the front-end electronics is the real bottleneck. Technological constraints limit the maximum number of independent electronics channels that can be brought to the peripheral electronics. Furthermore, the crosstalk between adjacent channels and the noise due to the high input capacitance to the preamplifiers become not negligible. In this case, it is the electronics chain that has to be brought to the individual pixel. We have implemented this concept by developing and building a CMOS VLSI array of 2101 pixels with 80 $\\mu$m pitch which is used directly as the charge collecting anode of the GEM. A description of the read-out ASIC for a MPGD and of its advantages is given in the next section. Section 3 describes the coupling of the chip die to the amplifying electrode, the assembly of the full detector and the results of laboratory tests obtained with a 5.9 keV X-ray source. The use of this new detection concept for Astronomical X-Ray Polarimetry and other applications are discussed in the last section. ", "conclusions": "A system in which the GEM foil, the absorption gap and the entrance window are assembled directly over a custom CMOS chip die has been developed. The transfer of charge from the amplifying region to the collection and read-out region occurs via electric fields. The ASIC itself becomes at the same time, the charge collecting anode and the pixelized read-out of a MicroPattern Gas Detector. For the first time the full electronics chain and the detector are completely integrated without the need of complicated bump-bonding. At a gain of 1000 a high sensitivity to single primary electron detection is reached. No problems have been found up to now in operating the system under HV and in a gas environment. An astronomical X-ray Polarimeter application has been presented. Final design will have 16$\\div$32 k channels and 60$\\div$70 microns pixel size ($\\simeq 1 cm^{2}$ active area). Depending on pixel and die size, electronics shaping time, analog vs. digital read-out, counting vs. integrating mode, gas filling, many others applications can be envisaged. This would open new directions in gas detector read-out, bringing the field to the same level of integration of solid state detectors." }, "0403/astro-ph0403162.txt": { "abstract": "We present a detailed analysis of a 3.5 s long burst from SGR1900+14 which occurred on 2001 July 2. The 2-150 keV time-integrated energy spectrum is well described by the sum of two blackbodies whose temperatures are approximately 4.3 and 9.8 keV. The time-resolved energy spectra are similarly well fit by the sum of two blackbodies. The higher temperature blackbody evolves with time in a manner consistent with a shrinking emitting surface. The interpretation of these results in the context of the magnetar model suggests that the two blackbody fit is an approximation of an absorbed, multi-temperature spectrum expected on theoretical grounds rather than a physical description of the emission. If this is indeed the case, our data provide further evidence for a strong magnetic field, and indicate that the entire neutron was radiating during most of the burst duration. ", "introduction": "\\label{intro} % KEVIN CAN YOU CHECK THE REFERRE's COMMENTS ON THE INTRO, AND ANSWER THEM... The soft gamma repeater \\sgr\\ was discovered in 1979 when it emitted 3 short bursts of soft gamma-rays in 3 days (Mazets et al. 1979). Its next recorded appearance occurred some 13 years later (Kouveliotou et al. 1993). Attempts were made over the years to obtain a precise position for the source and identify its counterpart (Hurley et al. 1994; Vasisht et al. 1994; Hurley et al. 1996), but this remained elusive until 1998, when the source entered a new period of activity, allowing it to be localized accurately by the interplanetary network (IPN: Hurley et al. 1999a). The precise source location was found to be consistent with that of a previously identified ROSAT quiescent X-ray source (Vasisht et al. 1994; Hurley et al. 1996). Observations with ASCA further revealed that the source had a 5.16 s period (Hurley et al. 1999b), and observations with RXTE demonstrated that the period was increasing rapidly ($\\rm 1.1 \\times 10^{-10}$ s s$^{-1}$, Kouveliotou et al. 1999). The counterpart to \\sgr~has not been found yet. If it is associated with the Galactic supernova remnant G42.8+0.6, it could be as close as 5 kpc (Hurley et al. 1999b). However, the source position also appears to be very close to a cluster of high mass stars, and it has been proposed that this may be the birthplace of the neutron star (Vrba et al. 2000). If so, its distance could be roughly 12-15 kpc. In all this paper we adopt a distance of 10 kpc. The 1998 activity of \\sgr~culminated in the giant flare of August 27 (Frail et al. 1999; Hurley et al. 1999c; Feroci et al. 1999; Mazets et al. 1999), which was followed by numerous, smaller bursts (e.g. Ibrahim et al. 2001). The next major period of activity of \\sgr~came in 2001 (Guidorzi et al. 2001; Hurley et al. 2001a, b; Feroci et al. 2001; Ricker et al. 2001a, b, c; Montanari et al. 2001, Feroci et al. 2004). During this episode, it became apparent that this source emits not only the common, short SGR bursts, with durations of about 200 ms and fluences $\\gtrsim 10^{-6} \\rm erg \\, cm^{-2}$, and the much rarer giant flares, lasting for minutes and having fluences $\\gtrsim 10^{-3} \\rm erg \\, cm^{-2}$, but also, a class of high-fluence \\it intermediate \\rm bursts, whose durations and fluences fall somewhere in between, and may in fact form a continuum (Kouveliotou et al. 2001; Woods et al. 2003; Feroci et al. 2003). Such events had been observed in the aftermath of the 1998 August 27 giant flare, but were thought to be related to it. In retrospect, it seems likely that such bursts were also emitted by SGR0525-66, in the aftermath of the famous 1979 March 5 giant flare (see Golenetskii et al. 1984). The properties of these intermediate bursts are interesting for numerous reasons, not the least of which are that some display X-ray afterglows, similar to the one observed during the 1998 August 27 event (Feroci et al. 2001; Thompson and Duncan et al. 2001), and that their longer duration permits studies of their spectral evolution. %OLD possess unique spectral properties. Duncan and Thompson (1992), Paczy\\'nski (1992), and Thompson and Duncan (1995, 1996) have proposed that the soft gamma repeaters are \\it magnetars \\rm, i.e., neutron stars with magnetic fields B $\\rm \\approx 10^{15} G$. In this model, magnetic dissipation causes the neutron star crust to fracture, and Alfv\\'en waves accelerate electrons, resulting in short (200 ms) bursts of soft gamma-radiation. Much more rarely, magnetic reconnection provides the energy for a longer, extremely energetic giant flare involving the entire neutron star magnetosphere. The periodicity observed in both the quiescent soft X-ray emission from \\sgr~and in the giant flare, the high spin-down rate, and the energetics of the giant flare, are all consistent with the main features of the magnetar model. In this paper, we analyze an intermediate burst from \\sgr , which occurred on 2001 July 2, emphasizing broad band and time-resolved X-ray spectral modeling obtained using the data of the FREGATE (FREnch GAmma-ray TElescope) and WXM (Wide Field X-ray Monitor) experiments aboard the HETE (High Energy Transient Explorer) spacecraft. ", "conclusions": "We have presented HETE-FREGATE and WXM observations of an intermediate burst from SGR1900+14. This is the first time that time-resolved spectral analysis of such a burst has been possible over a wide energy range, with good spectral and temporal resolution. We have found that the spectrum is well described by the sum of two blackbodies. One possible interpretation is that we are observing two distinct emitting volumes. Another is that we are observing emission which is affected by radiation transfer effects in a superstrong magnetic field. Both interpretations are consistent with the basic features of the magnetar model of soft gamma repeaters. However, it is not possible to choose between the two on the basis of a single observation. Fortunately, HETE-2 has observed numerous SGR bursts from both SGR1900+14 and SGR1806-20. Analysis of these observations is underway, and may provide less ambiguous support for one of these hypotheses. % Thus, this intermediate burst may % be viewed as a scaled-down giant flare. % JLA:SHOULD WE KEEP THAT ??? % JFO:Je crois pas." }, "0403/astro-ph0403118_arXiv.txt": { "abstract": "ESPaDOnS is a new-generation cross-dispersed \\'echelle spectropolarimeter, the commissioning phase of which is scheduled at the Canada-France-Hawaii Telescope (CFHT) for spring 2004. This instrument will provide full coverage of the optical domain (370~nm to 1,000~nm) in all polarization states (circular and linear) at a resolving power of about 70,000, with a peak efficiency of 20\\% (telescope and detector included). It includes a bench-mounted spectrograph, fiber-fed from a Cassegrain-mounted module including all polarimetric and calibration facilities. ESPaDOnS should be the most powerful tool dedicated to stellar spectropolarimetry, therefore opening unprecedented perspectives for major issues of stellar physics, from studies of stellar interiors to investigations of stellar atmospheres, stellar surfaces, stellar magnetic fields, and to observations of circumstellar environments and extra-solar planets. ", "introduction": "The combination of a polarimeter and a high-resolution \\'echelle spectrograph constitutes a powerful tool for investigating the physics of stellar atmospheres and environments. In particular, stellar magnetic fields are mostly detected by their polarizing effect on spectral lines. The analysis of this so-called ``Zeeman effect'' requires polarized, high resolution spectra, so that polarization across spectral line profiles can be investigated in detail, in order to measure both field strength and orientation. A spectral resolution in excess of 50,000 (to obtain at least two resolved elements in the thermal width of a typical spectral line) and the possibility to record linearly and circularly polarized spectra are therefore requested for optimal investigations of stellar magnetism. In addition to high resolution, a wide spectral coverage (preferentially covering the whole optical domain) is needed by spectroscopic studies (polarimetric or not) searching for tiny spectral signatures, the detection of which can greatly benefit from the rise in S/N produced by a simultaneous extraction of the signal in many spectral lines. To fulfill this range of ambitious technical performances, the ESPaDOnS concept was proposed by Donati et al. (1998). The instrument, developed by a collaboration including French, Canadian and Dutch experts, is now in its final assembling and testing phase at Observatoire Midi-Pyr\\'en\\'ees (France) before its first light scheduled at the CFHT during the second half of the year. We first describe the main characteristics of this instrument, which includes a polarimeter mounted on the Cassegrain bonnette of the telescope, fiber linked to a bench mounted spectrograph. We then evoke some of the most promising scientific issues that such a unique facility will make it possible to address in a near future. ", "conclusions": "" }, "0403/astro-ph0403432_arXiv.txt": { "abstract": "In February 1997 the Japanese radio astronomy satellite Halca was launched to provide the space-bourne element for the VSOP mission. Approximately twenty-five percent of the mission time has been dedicated to the VSOP Survey, a 5~GHz survey of bright, compact AGN. We present the results from the ongoing analysis. Both the final, calibrated, high resolution images and plots of visibility amplitude versus {\\em uv} distance for the first 102 of the sources have been prepared and has been submitted. Papers on the methods and the models from fitting the cumulative {\\em uv} amplitudes will also be submitted. The analysis of the second half is well underway. ", "introduction": "The radio astronomy satellite HALCA was launched by the former Institute of Space and Astronautical Science (now part of Japanese Aerospace eXpolaration Agency (JAXA))in February 1997 to participate in Very Long Baseline Interferometry (VLBI) observations with arrays of ground radio telescopes. It was was placed in an orbit with an apogee height above the Earth's surface of 21,400\\,km, a perigee height of 560\\,km, and an orbital period of 6.3~hours. HALCA provides the longest baselines of the VLBI Space Observatory Programme (VSOP), an international endeavor that has involved over 25 ground radio telescopes, five tracking stations and three correlators (Hirabayashi et al. 1998; 2000a). HALCA has now passed the end of the Guaranteed Observing Time period, and with the completion of the Memorandum of Understanding in February 2002 the NASA tracking and orbital calculation support ceased and the observation program has turned to completing the Survey. The orbital determination and spacecraft tracking are now completely indigenous to Japan and ISAS. ", "conclusions": "The satellite continues to make survey observations, and will do so while the satellite is functioning. These observations are being analysed cumulatively and individually and providing interesting results. These results are not only important for the understanding the target sources, mainly AGN's, but also for the planning of future space-VLBI missions such as the VSOP-2 mission, which will have a resolution of nearly a magnitude better." }, "0403/astro-ph0403604_arXiv.txt": { "abstract": "We evaluate the logarithmic derivative of the depth of the solar convective zone with respect to the logarithm of the radiative opacity, $\\partial \\ln R_{\\rm CZ}/\\partial \\ln \\kappa$. We use this expression to show that the radiative opacity near the base of the solar convective zone (CZ) must be known to an accuracy of $\\pm 1$\\% in order to calculate the CZ depth to the accuracy of the helioseismological measurement, $R_{\\rm CZ} = (0.713 \\pm 0.001)R_\\odot$. The radiative opacity near the base of the CZ that is obtained from OPAL tables must be increased by $\\sim 21$\\% in the Bahcall-Pinsonneault (2004) solar model if one wants to invoke opacity errors in order to reconcile recent solar heavy abundance determinations with the helioseismological measurement of $R_{\\rm CZ}$. We show that the radiative opacity near the base of the convective zone depends sensitively upon the assumed heavy element mass fraction, $Z$. The uncertainty in the measured value of $Z$ is currently the limiting factor in our ability to calculate the depth of the CZ. Different state-of-the-art interpolation schemes using the existing OPAL tables yield opacity values that differ by $\\sim 4$\\% . We describe the finer grid spacings that are necessary to interpolate the radiative opacity to $\\pm 1$\\%. Uncertainties due to the equation of state do not significantly affect the calculated depth of the convective zone. ", "introduction": "\\label{sec:intro} The depth of the solar convective zone has been measured by helioseismological techniques to high accuracy. In the most comprehensive study to date, Basu \\& Antia (1997)\\nocite{basucz} have investigated the influence of observational and theoretical systematic uncertainties as well as measurement errors. Basu and Antia concluded that the base of the solar convective zone currently lies at a depth of \\begin{equation} R_{\\rm CZ} ~=~ (0.713 \\pm 0.001)R_\\odot \\, . \\label{eq:rczmeasured} \\end{equation} The result of Basu and Antia is consistent with the earlier measurements of Kosovichev \\& Fedorova (1991),\\nocite{kosovichevcz} who obtained $R_{\\rm CZ} ~=~ (0.713 \\pm 0.001)R_\\odot$, and Christensen-Dalsgaard, Gough, \\& Thompson (1991),\\nocite{jcdCZ} who also obtained $R_{\\rm CZ} ~=~ (0.713 \\pm 0.003R_\\odot$, as well as with the determination of Guzik \\& Cox (1993),\\nocite{guzikcz} who found $R_{\\rm CZ} ~=~ (0.712 \\pm 0.001)R_\\odot$. Basu (1998)\\nocite{basu98} also studied the effect of the assumed value of the solar radius on the inferred depth of the convective zone and found $R_{\\rm CZ} ~=~ (0.7135 \\pm 0.0005)R_\\odot$. The analyses in these different studies span a wide range of reference solar models and analysis techniques. On the basis of the analyses cited above, the measurement of the depth of the solar convective zone appears robust and precise. Recently, new precision measurements have been made of the C, N, O, Ne, and Ar abundances on the surface of the Sun (Allende Prieto, Lambert, \\& Asplund 2001;\\nocite{allende01} Allende Prieto, Lambert, \\& Asplund 2002;\\nocite{allende02} Asplund et al. 2004;\\nocite{asplund04} Asplund et al. 2000;\\nocite{asplundetal00} Asplund 2000)\\nocite{asplund00}. These new abundance determinations use three-dimensional rather than one-dimensional atmospheric models, include hydrodynamical effects, and pay particular attention to uncertainties in atomic data and the observational spectra. The new abundance estimates, together with the previous best-estimates for other solar surface abundances~(Grevesse \\& Sauval 1998)\\nocite{oldcomp}, imply $Z/X = 0.0176$, much less than the previous value of $Z/X = 0.0229$ (Grevesse \\& Sauval 1998)\\nocite{oldcomp}. For a solar model with the recently-determined heavy element to hydrogen ratio, the calculated depth of the convective zone is~(Bahcall \\& Pinsonneault 2004)\\nocite{BP04} \\begin{equation} R_{\\rm CZ}(Z/X = 0.0176) ~=~ 0.726 R_\\odot \\, , \\label{eq:rcznewzoverx} \\end{equation} which is very different from the measured depth of the CZ (see equation~[\\ref{eq:rczmeasured}]). On the other hand, Basu and Antia (2004)~\\nocite{basu04} have shown that the helioseismological determination of $R_{\\rm CZ}$, equation \\ref{eq:rczmeasured}, is not affected if one assumes the correctness of the lower heavy element abundances ($Z/X = 0.0176$). Something is wrong. We have a new solar problem: ``the convective zone (CZ) problem.'' The radiative opacity is a key ingredient in calculating the depth of the convective zone. Moreover, about 95\\% of the total radiative opacity near the base of the convective zone involves bound electrons, either bound-free or bound-bound opacity (Iglesias 2004).\\nocite{iglesias04} Thus opacity calculations in this region involve details of the ionization balance and other delicate atomic physics properties. In this paper, we focus on determining the accuracy with which the opacity near the base of the convective zone must be known in order to calculate precisely the depth of the CZ with a stellar evolution code. We also evaluate the accuracy with which the opacity near the base of the CZ can be interpolated from OPAL tables. For a related comparison of the Los Alamos LEDCOP opacities and the OPAL opacities, see Neuforge-Verheecke et al. (2001)~\\nocite{losalamos}. For comprehensive discussions of stellar radiative opacities, the reader is referred to the important reviews by Rogers and Iglesias (1998)\\nocite{OPALreview} and by Seaton et al. (1995)\\nocite{opbook}. We investigate in a paper in preparation (Bahcall, Basu, Pinsonneault, and Serenelli 2004)~\\nocite{inpreparation} the helioseismological implications of the changes in opacity that are discussed in the present paper. The viability of any proposed change in the opacity discussed in the present paper must be tested by comparing a solar model that is evolved with the changed opacity with a complete set of precise helioseismological data. There is no compelling reason to believe that the illustrative change in opacity considered here, which is highly peaked in radius, will be either reproduced exactly by new opacity calculations or will be precisely consistent with helioseismological constraints. In the future, once new opacity calculations are available that satisfy the requirements described in this paper, it will be possible to test simultaneously the new opacities, the solar model evolution, and the helioseismological implications. We derive in \\S~\\ref{sec:dependence} the dependence, $\\partial \\ln R_{\\rm CZ}/\\partial \\ln \\kappa$, of the calculated depth of the solar convective zone upon the assumed radiative opacity. We apply this result to determine the accuracy with which the opacity must be known in order to calculate the depth of the CZ to the accuracy with which it is measured helioseismologically. We also determine the change in the standard OPAL opacity that is required in order to reconcile the helioseismological measurement with the recent determinations of heavy element abundances. We evaluate in \\S~\\ref{sec:composition} the dependence of the radiative opacity near the base of the convective zone upon the stellar composition. We find that the opacity depends sensitively upon the assumed heavy element abundance. We compare in \\S~\\ref{sec:comparison} the opacities obtained from two different interpolation schemes that are both applied to the same published OPAL opacity tables. Throughout this paper, we adopt the OPAL opacities (Iglesias \\& Rogers 1991a,b; Rogers \\& Iglesias 1992; Iglesias \\& Rogers 1996)\\nocite{iglesias91a,iglesias91b,rogers92,opalopacity96} as standard, when supplemented at low temperatures by values from Alexander \\& Fergusson (1994)\\nocite{alexanderopac}. We use simulated opacity tables in \\S~\\ref{sec:shifted} to estimate the likely uncertainties that result from interpolations within the existing OPAL opacity tables and to determine the grid sizes to obtain small interpolation errors. For completeness and for contrast, we use four different equations of state to show in Appendix~A that uncertainties due to the choice of EOS are not important, at the present level of accuracy, for the calculation of the depth of the solar convective zone. We also demonstrate in Appendix~B that uncertainties in the nuclear reaction rates affect the depth of the solar convective zone only at the level of 0.1\\% . In Appendix~C, we verify that the conversion of carbon and oxygen in CNO burning, which cannot be accurately modeled with existing opacity codes, causes a 0.1\\% uncertainty in the calculated depth of the convective zone. Basu and Antia (2004) (see also Asplund et al. 2004) have shown that errors in the calculation of the diffusion coefficients are unlikely to be the correct explanation of the discrepancy between measured and calculated depth of the solar convective zone. Other solar model ingredients, including the element diffusion coefficients, can affect the calculated depth of the convective zone. A complete investigation of all the possible effects on the convective zone is beyond the scope of the present paper and would distract the reader from our main concern, the effect of the radiative opacity. Moreover, we believe that the radiative opacity and the heavy element abundance provide the single largest contributions to the error budget for the calculation of the solar convective zone. The effect of the heavy element abundance on the calculated depth of the convective zone has been evaluated in Bahcall and Pinsonneault (2004). We summarize and discuss our main results \\hbox{in \\S~\\ref{sec:summary}}. ", "conclusions": "\\label{sec:summary} The primary goal of this paper is to determine how accurately the radiative opacity near the base of the convective zone must be known in order to use measurements of the CZ depth to draw conclusions about other solar parameters. There are two separate but related issues with respect to the accuracy of the radiative opacity, namely, the accuracy with which the tabulated values in opacity tables are calculated and the accuracy with which the opacity can be interpolated within tables of a specified grid size. We first summarize our conclusions regarding the accuracy of tabulated opacity values and then we summarize our results with respect to the accuracy of interpolations within the standard OPAL opacity tables. The helioseismological implications of the opacity changes considered in this paper will be discussed in Bahcall, Basu, Pinsonneault, and Serenelli, (2004, in preparation). We show in \\S~\\ref{sec:dependence} that the logarithmic derivative of the convective zone depth with respect to the logarithm of the opacity satisfies $\\partial \\ln R_{\\rm CZ}/\\partial \\ln \\kappa \\approx -0.095$. We conclude from this relation that the radiative opacity must be known to an accuracy of 1\\% in order to calculate in a solar model the depth of the CZ to the accuracy, 0.14\\%, with which the depth is measured by helioseismology. On the other hand, if one accepts the recent measurements of heavy element abundances, then the OPAL opacities must be increased by about 21\\% in order to reconcile the calculated solar model depth of the CZ and the measured depth of the CZ. This change of 21\\% could conceivably arise from a combination of errors in the tabulated values of the opacity and interpolation errors, which are discussed below. However, as we shall see, the total change of 21\\% is too large to be ascribed solely to errors in interpolation. It would be very instructive to have a comprehensive study of the absolute accuracy of state-of-the-art radiative opacity calculations. A detailed comparison of the calculated opacity near the base of the convective zone obtained by the Opacity Project (Seaton, Yan, Mihalis, \\& Pradhan 1994)~\\nocite{opacityproject} with the results of the OPAL project (Iglesias \\& Rogers 1996) would be very informative. The interested reader is referred to the informative and insightful comparison by Neuforge-Verheecke et al. (2001) of the Los Alamos LEDCOP opacities and the OPAL opacities. The largest differences are found near the base of the convective zone, with the OPAL opacities being as much as 6\\% larger than the LEDCOP opacities in this region. As part of a comprehensive discussion of factors that affect the accuracy of solar models, Boothroyd \\& Sackmann (2003)~\\nocite{sackmann03} have investigated ways that the opacities can affect helioseismological parameters. We show in \\S~\\ref{sec:composition} that the radiative opacity near the base of the convective zone depends sensitively upon the assumed chemical composition (see especially equation~\\ref{eq:kappaoncomposition} and equation~\\ref{eq:compositionderivatives}). If one wanted to calculate the depth to an accuracy of $0.6$\\%, then one would need to know the heavy element mass fraction, $Z$, to an accuracy of 1\\%. This precision is far beyond the current state-of-the-art accuracy in the determination of the heavy element abundance. The entire difference between the measured depth of the solar convective zone (equation~\\ref{eq:rczmeasured}) and the value calculated using a solar model with the recent low determinations of the heavy element abundances (equation~\\ref{eq:rcznewzoverx}) could be explained by the present uncertainty, $\\sim 15$\\%, in the ratio of Z/X (see Bahcall \\& Pinsonneault 2004)\\nocite{BP04}. Of course, the changes in opacity caused by changing $Z/X$ are not limited to any particular region. Changing the assumed surface value of $Z/X$ affects the composition and hence the opacity throughout the solar model. We have approximated in this paper the dependence of the opacity upon composition by the dependence upon just two variables, the mass fractions $X$ and $Z$. In reality, the situation is more complex. Different chemical elements contribute differently to the stellar opacity. For example, Bahcall, Pinsonneault, and Basu (2001) found that the depth of the convective zone was most sensitive to the abundances of the lighter metals, which are significant opacity sources at $2 \\times 10^6$K, while the heavier metals were much more important for the core structure and the estimated initial solar helium abundance. However, we are not yet at a level of precision that we can specify well the opacity-weighted uncertainties of the different heavy elements. This is a refinement that will have to await further progress in determining the different heavy element abundances and more extensive opacity calculations. We compare in \\S~\\ref{sec:comparison} the radiative opacity values obtained with two different interpolation routines from the standard OPAL opacity tables. We find that the difference in interpolated values of the radiative opacity can be as large as 4\\% near the base of the convective zone. We also tested in \\S~\\ref{sec:shifted} the accuracy with which interpolations can be performed within simulated opacity tables of different grid sizes. We find that errors of the order of 3\\% may be expected from tables with the grid spacings of the existing OPAL tables. However, we show that the interpolation uncertainties could be reduced to the level of 1\\% or below by using a denser grid with $\\Delta \\log T = 0.025$, $\\Delta \\log r = 0.125$, and with $Z$ ranging from $Z = 0.0100$ to $Z = 0.0225$ with $\\Delta Z = 0.0025$. For completeness, we report in the Appendix on the calculated depth of the CZ that was found using four different equations of state. In agreement with other authors, we find that the choice of equation of state affects the calculated depth of the CZ by only about $\\pm 0.1$ \\%. We also show in the Appendix that current uncertainties in nuclear reaction rates also affect the calculated depth of the convective zone at the level of 0.1\\%." }, "0403/astro-ph0403097_arXiv.txt": { "abstract": "Milagro is a water Cherenkov extensive air shower array that continuously monitors the entire overhead sky in the TeV energy band. The results from an analysis of $\\sim$3 years of data (December 2000 through November 2003) are presented. The data has been searched for steady point sources of TeV gamma rays between declinations of 1.1 degrees and 80 degrees. Two sources are detected, the Crab Nebula and the active galaxy Mrk 421. For the remainder of the Northern hemisphere we set 95\\% C.L. upper limits between 275 and 600 mCrab (4.8-10.5$\\times 10^{-12}$ cm$^{-2}$ s$^{-1}$) above 1 TeV for source declinations between 5 degrees and 70 degrees. Since the sensitivity of Milagro depends upon the spectrum of the source at the top of the atmosphere, the dependence of the limits on the spectrum of a candidate source is presented. Because high-energy gamma rays from extragalactic sources are absorbed by interactions with the extragalactic background light the dependence of the flux limits on the redshift of a candidate source are given. The upper limits presented here are over an order of magnitude more stringent than previously published limits from TeV gamma-ray all-sky surveys. ", "introduction": "\\label{sec:Introduction} Sources of very-high-energy (VHE, $>$100 GeV) gamma rays are observed to be non-thermal in nature and are typically the sites of particle acceleration. This acceleration is thought to occur in astrophysical shocks such as those believed to exist in plerions \\citep{deJager1992}, supernova remnants \\citep{Volk2000}, active galactic nuclei \\citep{Blandford1978}, and galaxy clusters \\citep{Loeb2000} (among other sources). These shocks may accelerate protons or electrons, both of which lead to the emission of gamma rays. Since gamma rays are unaffected by the magnetic fields that pervade the Galaxy and the Universe, they can be used to pinpoint the sites of particle acceleration. In addition to these ``classical'' astrophysical sources of VHE gamma rays, other more exotic objects such as primordial black holes, topological defects, and the decay of relic particles from the big bang may also emit VHE gamma rays. Perhaps of most interest is the possible existence of a new type of source that has yet to be postulated. A comprehensive survey of the sky sensitive to emission at all time scales is necessary to detect the many possible sources. The analysis presented in this paper is part of an ongoing effort by the Milagro collaboration to search the entire northern hemisphere for such objects. The search for short bursts of TeV gamma rays has been addressed in a previous paper \\citep{Atkins2004}. The analysis presented here deals specifically with steady point sources of TeV gamma rays. It has been over a decade since the discovery of the first source of VHE gamma rays, the Crab Nebula \\citep{Weekes89}. Since that time there have been seven other confirmed sources of TeV gamma rays \\citep{Horan2003}, six of which lie in the northern hemisphere. With the exception of the Crab Nebula and the supernova remnant PKS 1706-44, these objects are all active galactic nuclei of the blazar class \\citep{Horan2003}. All of these objects have been discovered by atmospheric Cherenkov telescopes searching for counterparts to sources discovered at lower energies. In contrast, the EGRET instrument detected over 270 objects emitting high-energy gamma rays above 100 MeV \\citep{Hartman99}. One hundred and seventy of these objects are not identified at other wavelengths. The VHE regime is a natural energy band to search for counterparts to these objects. The small field-of-view and low duty factor of the atmospheric Cherenkov telescopes (ACTs) make comprehensive sky surveys difficult to perform. As a result, very few comprehensive surveys of the VHE sky have been performed to date. The first VHE survey was performed by a non-imaging ACT \\citep{Helmken79, Weekes79} in 1979. The Milagrito instrument (a prototype to Milagro, with no background rejection capability and a higher energy threshold) also performed a survey of the northern hemisphere \\citep{Atkins2001} and set limits of $\\le$3 Crab from any point source in the northern hemisphere. More recently flux limits of 4-9 times that of the Crab Nebula (above 15 TeV) have been published by the AIROBICC collaboration \\citep{Aharonian2002}. (1 Crab is equivalent to an integral flux above 1 TeV of $F(>$1 TeV)$=1.75 \\times$10$^{-11}$ cm$^{-2}$ s$^{-1}$.) The limits presented here are over an order of magnitude more stringent than these previous surveys. Since many of the confirmed sources of VHE gamma rays are extragalactic the limits must account for the absorption of TeV gamma rays by interactions with the extragalactic background light (EBL) \\citep{Primack2000, Stecker2002, Kneiske2002}. The EBL is comprised of visible radiation emitted by stars and infrared radiation emitted by dust due to reprocessed starlight. Since direct measurements of the intensity and spectrum of the EBL are problematic due to the foreground light from our galaxy, a model is used to determine the effect of the EBL on the observed spectrum at earth from a distant source. Before employing the results presented here the limitations of this survey need to be understood. First, the limits presented only apply to point sources, not to extended objects, such as the galactic plane. Second, they only apply to the average VHE emission during the time period over which the data was obtained. ", "conclusions": "\\label{sec:Conclusions} A complete survey of the northern hemisphere (declination $1.1^{\\circ}$ to $80^{\\circ}$) for point sources of TeV gamma rays has been performed. These limits apply to the average flux level during the roughly 3 year period from December 15, 2000 through November 25, 2003. The average 95\\% C.L. upper limits range from 275 mCrab to 600 mCrab depending upon the declination of the source and are over an order of magnitude more restrictive than previous limits. A prescription has been given to calculate the corresponding upper limits for sources with different spectra and for extragalactic sources. For sources with differential spectral indices of -2.0 the upper limits are 57\\% lower. For a source at a redshift of 0.03 the flux limits are a factor of 2.4 larger. While these limits are the best available to date, Milagro has recently been completed with the construction of an array of 175 water tanks surrounding the central reservoir. A comparable dataset, with this now complete Milagro detector, would improve these limits by a factor $\\sim$2." }, "0403/astro-ph0403574_arXiv.txt": { "abstract": " ", "introduction": "The current paradigm for the formation of giant gaseous planets is based on the so-called core accretion model in which a growing solid core reaches a critical mass and accretes rapidly a massive atmosphere (Pollack et al. \\cite{P96}, hereafter referred to as P96). While this model has many appealing features, it suffers at least from two shortcomings which, as we shall show later, are actually coupled. First, the timescale (close to 10 Myr) found by P96 to form Jupiter at its present location is uncomfortably close to the typical lifetime of protoplanetary discs which is believed to be of the order of 1-10 Myr (Haisch et al. \\cite{Haisch}). This timescale problem has led others to look for more rapid formation mechanisms based on direct gravitational collapse (Boss \\cite{boss02}). Second, P96 assumed that the giant planets of our solar system have been formed where they are observed today. However, the discovery over the last decade of extrasolar planets at very short distances to their parent star, has open the possibility that planets may actually migrate over large distances (Lin et al. \\cite{linetal96}; Trilling et al. \\cite{Trilling}; Papaloizou \\& Terquem \\cite{PT99}, hereafter referred to as PT99). The timescale of migration is still very uncertain, but conservative estimates give values between $0.1$ and $10$ Myr. This timescale is therefore comparable to the planet formation timescale and the disc lifetime and thus migration cannot be neglected in a self-consistent picture. In this Letter, we extend the core accretion model of P96 to include migration, disc evolution and gap formation and show that the formation timescales of giant planets can be reduced by factors of ten or more. ", "conclusions": "We have calculated the formation of giant planets up to runaway gas accretion and studied the effect of migration and gap formation on the resulting formation timescale. Our main result is that phase 2 described in P96 is suppressed, and this leads to a formation time around $1$ Myr for the initial conditions chosen here. This is roughly an order of magnitude faster that the favored model of P96 even though our initial disc had less mass than the one used by P96. We have performed many tests and convinced ourselves that this speed-up is robust against changes in various parameters. For example, in a calculation in which the reduction of type I migration ($f_I$) is set to 0.1, an embryo starting at $15$AU undergoes runaway accretion in less than 3 Myr. The assumed size of the planetesimals plays a critical role as already noted by P96. For example, assuming planetesimals of 10km instead of 100km leads to runaway accretion after only 0.3 Myr! As already mentioned in section \\ref{infalling}, since the effect of ablation is negligible in our calculations, mass loss of planetesimals occurs very deep in the envelope. The switch from sinking to no-sinking approximation (see P96) has then a small effect. In the migrating case (without gap formation) we obtain a formation timescale of $\\sim 1.2$ Myr using the sinking approximation, compared to $\\sim 1$ Myr in the no-sinking case. Thus, migration, besides explaining the presence of giant planets at short distances to their stars, also plays an important role in the formation process itself. By ensuring that the feeding zone is never depleted, migration suppresses almost entirely the protracted phase 2 therefore shortening enormously the formation time. This of course, does not preclude other effects such as reduction of opacity or formation of vortices prior to planet formation to further reduce this timescale. The formation of giant planets through the core-accretion scenario can therefore proceed over timescales in good agreement with disk lifetimes, and {\\it without having to consider disks significantly more massive than the minimum mass solar nebula}. Finally, we point out that the formation of giant planets appears only possible if the currently available type I migration rates are reduced by at least a factor of 10. This suggests that their might still be a serious problem in our understanding of this type of migration. The authors would like to acknowledge the help of C. Winisdoerffer and the ENS-Lyon team of the C.R.A.L. We thank P. Bodenheimer and G. Meynet for useful discussions. This work was supported in part by the Swiss National Science Foundation." }, "0403/astro-ph0403268_arXiv.txt": { "abstract": "Current data exclude cosmic strings as the primary source of primordial density fluctuations. However, in a wide class of inflationary models, strings can form at later stages of inflation and have potentially detectable observational signatures. We study the constraints from WMAP and SDSS data on the fraction of primordial fluctuations sourced by local cosmic strings. The Bayesian analysis presented in this brief report is restricted to the minimal number of parameters. Yet it is useful for two reasons. It confirms the results of \\cite{PTWW03} using an alternative statistical method. Secondly, it justifies the more costly multi-parameter analysis. Already, varying only three parameters -- the spectral index and the amplitudes of the adiabatic and string contributions -- we find that the upper bound on the cosmic string contribution is of order $10\\%$. We expect that the full multi-parameter study, currently underway \\cite{PWW04}, will likely loosen this bound. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403560_arXiv.txt": { "abstract": "We construct eigenspectra from the DR1 quasars in the SDSS using the Karhunen-Lo\\`eve (KL) transform (or Principal Component Analysis, PCA) in different redshift and luminosity bins. We find that the quasar spectra can be classified, by the first two eigenspectra, into a continuous sequence in the variation of the spectral slope. We also find a dependence on redshift and luminosity in the eigencoefficients. The dominant redshift effect is the evolution of the blended Fe~II emission (optical) and the Balmer continuum (the ``small-bump'', $\\lambda_{rest} \\approx 2000-4000$\\AA), while the luminosity effect is related to the Baldwin effect. Correlations among several major broad emission lines are found, including the well-known ``Eigenvector-1''. ", "introduction": "Spectroscopic observations from the Sloan Digital Sky Survey (SDSS) have the advantage of large numbers of quasars (QSOs) and a large redshift range that provides an unique opportunity to study their intrinsic properties, sample variation, and spectral classification in great detail. A powerful method to address these questions is the KL transform. The idea is to derive from the observed spectra a lower dimensional set of eigenspectra (Connolly et al.\\ 1995), in which the essential spectral properties are represented. In the following we outline some results from applying this method to the SDSS quasars. ", "conclusions": "" }, "0403/astro-ph0403083_arXiv.txt": { "abstract": "Several short-period Jupiter-mass planets have been discovered around nearby solar-type stars. During the circularization of their orbits, the dissipation of tidal disturbance by their host stars heats the interior and inflates the sizes of these planets. Based on a series of internal structure calculations for giant planets, we examine the physical processes which determine their luminosity-radius relation. In the gaseous envelope of these planets, efficient convection enforces a nearly adiabatic stratification. During their gravitational contraction, the planets' radii are determined, through the condition of a quasi-hydrostatic equilibrium, by their central pressure. In interiors of mature, compact, distant planets, such as Jupiter, degeneracy pressure and the non-ideal equation of state determine their structure. But, in order for young or intensely heated gas giant planets to attain quasi-hydrostatic equilibria, with sizes comparable to or larger than two Jupiter radii, their interiors must have sufficiently high temperature and low density such that degeneracy effects are relatively weak. Consequently, the effective polytropic index monotonically increases whereas the central temperature increases and then decreases with the planets' size. These effects, along with a temperature-sensitive opacity for the radiative surface layers of giant planets, cause the power index of the luminosity's dependence on radius to decrease with increasing radius. For planets larger than twice Jupiter's radius, this index is sufficiently small that they become unstable to tidal inflation. We make comparisons between cases of uniform heating and cases in which the heating is concentrated in various locations within the giant planet. Based on these results we suggest that accurate measurement of the sizes of close-in young Jupiters can be used to probe their internal structure under the influence of tidal heating. ", "introduction": "One of the surprising findings in the search for planetary systems around other stars is the discovery of extrasolar planets with periods down to 3 days (\\cite{mq95}). Nearly all planets with period less than 7 days have nearly circular orbits. In contrast, known extrasolar planets with periods longer than 2-3 weeks, have nearly a uniform eccentricity distribution. The shortest-period planets and their host stars induce tidal perturbations on each other. When these disturbances are dissipated, angular momentum is exchanged between the planets and their host stars, leading toward both a spin synchronization and orbital circularization (\\cite{ra96}). \\cite{BLM} (hereafter Paper I), considered the effect of tidal dissipation of energy during the synchronization of these planets' spin and the circularization of their orbits. In that analysis, they compute a series of numerical models for the interior structures of weakly eccentric Jovian planets at constant orbital distances under the influence of interior tidal heating and stellar irradiation. In these previous calculations, the interior heating rate per unit mass was imposed to be constant in time and uniformly distributed within the planet. Under these assumptions, they showed that Jovian planets can be inflated to equilibrium sizes considerably larger than those deduced for gravitationally contracting and externally heated planets. For the transiting planet around HD 209458, they suggested that provided its dimensionless dissipation $Q$-value is comparable to that inferred for Jupiter (\\cite{yp81}), a small eccentricity ($e \\simeq 0.03$) would provide adequate tidal heating to inflate it to its observed size (\\cite{brown01}). Since the orbital circulation time scale is expected to be shorter than the life span of the planet, they also suggested that this eccentricity may be excited by another planet with a longer orbital period. The prediction of a small eccentricity and the existence of another planet are consistent with existing data (\\cite{bll03}). There are at least two other scenarios for the unexpectedly large size of HD 209458b. Heating by stellar irradiation reduces the temperature gradient and the radiative flux in the outer layers of short-period planets. This process could significantly slow down the Kelvin-Helmholtz contraction of the planet and explain the large size (\\cite{bur00}). However, even though the stellar flux onto the planet's surface is 5 orders of magnitude larger than that released by the gravitational contraction and cooling of its envelope, this heating effect alone increases the radius of the planet by about 10\\%, not by 40\\% as observed (\\cite{gs02}). An alternative source is the kinetic heating induced by the dissipation of the gas flow in the atmosphere which occurs because of the pressure gradient between the day and night sides (\\cite{gs02}). In order to account for the observed size of the planet, conversion of only 1\\% of the incident radiative flux may be needed, provided that the dissipation of induced kinetic energy into heat occurs at sufficiently deep layers (tens to 100 bars). \\cite{sg02} suggest that the Coriolis force associated with a synchronously spinning planet may induce the circulation to penetrate that far into the planet's interior, and that dissipation could occur through, for example, Kelvin-Helmholtz instability. A follow-up analysis suggests that this effect may be limited (\\cite{blb03}; \\cite{jl03}). In order to distinguish between these three scenarios, the effect of tidal heating for planets with modest to large eccentricities was considered. In a follow-up paper (Gu {\\it et al.} 2003, hereafter Paper II), we showed that the size ($R_p$) of compact Jupiter-mass planets slowly increases with the tidal dissipation rate. For computational simplicity, we adopted a conventional equilibrium tidal model which describes the planets' continuous structural adjustment in order for them to maintain a state of quasi-hydrostatic equilibrium in the varying gravitational potential of their orbital companion. In this prescription, a phase lag into the response is introduced to represent the lag being proportional to the tidal forcing frequency and attributable to the viscosity of the body. The phase lag gives rise to a net tidal torque and dissipation of energy and the efficiency of the tidal dissipation can be parametrized, whatever its origin, by a specific dissipation function or $Q$-value (quality factor) (\\cite{gs66}). External perturbation can also induce dynamical tidal responses through the excitation of g modes (Ioannou \\& Lindzen 1993a,b) or inertial waves (\\cite{ol04}) which can be damped by viscous dissipation in the interior (\\cite{gn77}) or radiative and nonlinear dissipation in the atmosphere of the planet (\\cite{ltl97}). In both the prescription for equilibrium and dynamical responses, the tidal dissipation rate is a rapidly increasing function of the planets' radius. But, their surface luminosity increases even faster with $R_p$ such that, planets with relatively small eccentricities and modest to long periods attain a state of thermal equilibrium in which the radiative loss on their surface is balanced by the tidal dissipation in their interior. For planets with short periods and modest to large eccentricities, the rate of interior heating is sufficiently large that their $R_p$ may inflate to more than two Jupiter radii. In this limit, the surface luminosity of the planet becomes a less sensitive function of its $R_p$ and the eccentricity damping rate is smaller than the expansion rate of the planet so that the increases in their surface cooling rate cannot keep pace with the enhanced dissipation rate due to their inflated sizes. These planets are expected to undergo runaway inflation and mass loss. We suggested that the absence of ultra-short-period Jupiter-mass planets with $P<3$ days, which corresponds to an orbital semi-major axis $a$ of 0.04 AU, may be due to mass loss through Roche-lobe overflow resulting from such a tidal inflation instability. Other scenarios have been proposed to explain the lack of Jupiter-mass planets with $P<3$ days: truncation of inner part of a disk (\\cite{lin96}; \\cite{kl02}), orbital migration due to the spin-orbit tidal interaction between the close-in planet and the parent star (\\cite{ra96}; \\cite{ws02}; \\cite{pr02}; \\cite{jiang03}), and Roche-overflowing planets with the help of disk-planet interaction but without tidal inflation (\\cite{tri98}). In this contribution, we continue our investigation on the internal structure of tidally heated short-period planets. The main issues to be examined here are: 1) how does tidal energy dissipation actually lead to the expansion of the envelope? 2) how does the internal structure of the planet depend on the distribution of their internal tidal dissipation rate? and 3) what are the important physical effects which determine the tidal inflation stability of the planets? These issues are important in determining the mass-radius relation of short-period planets which is directly observable. Structural adjustments may also modify the efficiency of dissipation and the planets' Q-value for both equilibrium and dynamical tides. In the extended convective envelopes of gaseous giant planets and low-mass stars, turbulence can lead to dissipation of the motion that results from the continual adjustment of the equilibrium tide. However, the turbulent viscosity estimated from the mixing-length theory ought to be reduced by a frequency-dependent factor owing to the fact that the convective turnover time scale is usually much longer than the period of the tidal forcing. Based on the present-day structure of Jupiter, Goldreich \\& Nicholson (1977) estimated $Q\\approx5\\times10^{13}$. However, within the intensely heated (by tidal dissipation) short-period extra solar planets, convection is expected to be more rigorous with higher frequencies whereas their tidal forcing frequencies are smaller than that of Jupiter. The Q-value for the equilibrium tide within extra solar planets is likely to much smaller than that within Jupiter. For the dynamical response of short-period extra solar planets planet, the forcing frequencies are typically comparable to their spin frequencies but are small compared to their dynamical frequencies. Convective regions of the planets support inertial waves, which possess a dense or continuous frequency spectrum in the absence of viscosity, while any radiative regions support generalized Hough waves (Ioannou \\& Lindzen 1993a, b). Inertial waves provide a natural avenue for efficient tidal dissipation in most cases of interest. The resulting value of $Q$ depends, in principle, in a highly erratic way on the forcing frequency. Since the planets' spin frequency adjusts with their $R_p$, which is a time-delayed function of the tidal dissipation rate within them, the efficiency of tidal dissipation may fluctuate while the overall evolution is determined by a frequency-average Q-value (\\cite{ol04}). In \\S2, we briefly recapitulate the basic equations which determine the quasi-static evolution of the planets' structure. We show the simulation results for inflated giant planets in the case of constant internal heating per unit mass in \\S3 and analyze the results, in the Appendix, in terms of polytropic models which allow us to conclude that the onset of the tidal runaway inflation instability is regulated by a transition in the equation of state for the interior gas from a partially degenerate/non-ideal state toward a more ideal-gas state. We examine the dependence of planetary adjustment on different locations of the energy dissipation in \\S4. Finally, we summarize the results and discuss their observational implications in \\S5. ", "conclusions": "In this paper, we continue our investigation on the adjustment of a planetary interior as a consequence of intense tidal heating. As a giant planet's interior is heated and inflated, we showed in Paper II that its interior remains mostly convective. Efficient energy transport leads to an adiabatic stratification. With a constant heating rate per unit mass, we deduced an unique luminosity-radius relation regardless of how intense the heating rate is. The planet's luminosity increases with its radius. But the growth rate of $\\cal L$ is a decreasing function of $R_p$. At the same time, the tidal dissipation heats the interior of the planet at a rate which increases rapidly with $R_p$. At around $2~R_J$, $\\cal L$ can no longer sustain sufficient growth to maintain a thermal equilibrium with the tidal dissipation rate. Thereafter, the planet's inflation become unstable and it overflows its Roche radius and become tidally disrupted. Here we show that the change of luminosity during the planet's expansion is directly linked to the evolution of its interior, in particular, the equation of state. We employ the polytrope approach to investigate the interior structure of an inflated giant planet. According to the simulation, interior profiles deviate away from $P\\propto \\rho^2$ as the planet expands. The central temperature $T_c$ increases and then decreases with the size of the planet. Also $P_c$ and $\\rho_c$ are less steep functions of $R_p$ than the polytrope theory with a constant polytropic index $n$ indicates. All of these effects suggest that the planet interior does not evolve in a self-similar manner, but $n$ gets larger as $R_p$ increases. In conjunction with the numerical results that the degeneracy $D$ decreases, and that $T_c$ rises and then drops during the course of inflation, the result of a positive value of $dn/dR_p$ can be interpreted as a manifestation of a reduction in degeneracy during the expansion. We reason that the coefficient of thermal expansion $\\delta$ increases in response to a decrease in degeneracy and non-ideal effects, leading to an increase in $n$ through equation (\\ref{eq:n_delta}). Consequently the planet compressibility at constant mass $C_M$ increases with $R_p$ (see equation \\ref{eq:C_M}). This pattern can be translated into the phenomenon of a decrease in luminosity growth during the inflation, as a consequence of an one-to-one relation between $K$ and $\\cal L$ in the case of uniform heating in mass under the conditions of the polytropic interior and hydrostatic equilibrium. We also compare the results between a planet with a core and without a core. To be inflated to the same size, a planet with a core, therefore possessing a larger gravitational binding energy, needs a larger intrinsic luminosity $\\cal L$ than a planet with no core and the same mass. We also show that the opacity in the radiative envelope has a drastic effect on the final equilibrium size of an inflated planet: the size would be much smaller if grains are depleted in the radiative envelope. We also consider the possibility of localized tidal dissipation. Such a process may occur in differentially rotating planets or near the interface between the convective and radiative zones where the wavelength associated with dynamical tidal response is comparable to the density scale height. Localized dissipation may also occur through the dissipation of resonant inertial waves or radiative damping in the atmosphere. In the strong shell-heating models, the one-to-one relation between $K$ and $\\cal L$ disappears because of the existence of a radiative region caused by temperature inversion beneath the shell-heating zone. The unheated planet's interior in such cases might still be inflated due to a significant expansion of the gas above the heating zone, although the overall expansion rate is less efficient than that in the uniform heating cases as a result of a greater amount of radiative loss from the planet's photosphere. Without gaining entropy, the expanding interior cannot lift its degeneracy and therefore cannot increase its elasticity. However, the gas above the shell-heating zone can lift its non-ideal properties and hence enhance its elasticity, leading to a decrease in $\\Gamma$ in that region and thereby diminishing the growth rate of $\\cal L$ as the planet expands. Finally, we consider the self consistent response, taking into account the modification of heating rate due to a planet's expansion. In this paper, we adopt a constant-$Q$ prescription for equilibrium tides in which the tidal dissipation rate is assumed to be proportional to $R_p^5$. The results for the uniform heating model suggest that a young gaseous planet of $1~M_J$ without a solid core can be thermally unstable and inflated from $2~R_J$ to a size beyond $4~R_J$ if $e\\sim 0.2$ at $a=10~\\rsun$. If the dissipation rate is proportional to $R_p^5$, and if most of the tidal perturbation is deposited at $m/M_p=0.9$, a core-less young planet of $1 M_J$ would be thermally unstable and inflated from $2 R_J$ to a size beyond $3 R_J$ for $e>0.07$ at $a=0.03$ AU or $e>0.26$ at $a=10 \\rsun$. With the same heating concentration $m/M_p=0.9$ and $R_p$-dependence in the dissipation rate, a young planet with a core at $a=0.03$ AU with an initial eccentricity $e>0.294$ can be inflated from $2~R_J$ to a size beyond its Roche radius. We have assumed that the convective flow still behaves adiabatically even though the heating shell causes a narrow radiative zone. However, the condition away from adiabaticity implies that the internal heat is not transported away as efficiently as in the case of adiabatic convection, leading to a more severe reduction in non-ideal properties of the gas and therefore an even faster decrease of $\\gamma$ as the planet expands. Note that the Eddington approximation for the surface boundary condition is used in these models rather than more detailed frequency-dependent model atmospheres. This approximation is not necessarily valid for the strongly irradiated atmospheres studied here (\\cite{gs02}). However it is unlikely to make much difference for the main results discussed here, namely the behavior of the planet's radius as a function of tidal dissipation energy. It could, however, lead to errors in other kinds of predictions, such as the radius as a function of opacity. The equilibrium tidal dissipation formula is based on an {\\it ad hoc} assumption of a constant lag angle. In reality, the dynamical tidal response of a planet through both gravity and inertial waves near the planet's surface and convective envelope may be much more intense, especially through global normal modes. Their dissipation may provide the dominant angular momentum transfer mechanism for the orbital evolution and heating sources for the internal structure of close-in extrasolar planets. In the limit of small viscosity, the intensity of tidal dissipation is highly frequency dependent (Ogilvie \\& Lin 2004). When the forcing and response frequencies are in resonance, the energy dissipation rate is intense whereas between the resonances it is negligible. As the planets undergo structure adjustments, their spin frequency, Brunt--V\\\"ais\\\"al\\\"a frequency distribution, the adiabatic index, and equation of state also evolve. Since all of these physical effects contribute to the planets' dynamical response to the tidal perturbation from their host stars, their response and resonant frequencies are continually modified. The results in this paper indicate that the structure of the planet adjusts on a radiation transfer time scale which generally differs from the time scale for a planet to evolve through the non resonant region. In addition, the tidal forcing frequency also changes as the planets evolve toward a state of synchronous spins and circular orbits. Therefore, it is more appropriate to consider a frequency averaged tidal dissipation rate. In the limit of small viscosity, the frequency averaged dissipation rate converges (Ogilvie \\& Lin 2004) such that the equilibrium tidal dissipation formula may be a reasonable approximation. Nevertheless, we cannot yet rule out the possibility that some close-in planets may attain some non resonant configuration and stall their orbital evolution. Therefore, accurate measurement of the sizes of close-in young Jupiters via planet transit surveys can be used to constrain the theories of tidal dissipation and hence internal structure for these objects." }, "0403/astro-ph0403610_arXiv.txt": { "abstract": "The mass discrepancy in disk galaxies is shown to be well correlated with acceleration, increasing systematically with decreasing acceleration below a critical scale $a_0 \\approx 3700\\kmskpc = 1.2 \\times 10^{-10}\\;\\mathrm{m}\\,\\mathrm{s}^{-2}$. For each galaxy, there is an optimal choice of stellar mass-to-light ratio which minimizes the scatter in this mass discrepancy-acceleration relation. The same mass-to-light ratios also minimize the scatter in the baryonic Tully-Fisher relation and are in excellent agreement with the expectations of stellar population synthesis. Once the disk mass is determined in this fashion, the dark matter distribution is specified. The circular velocity attributable to the dark matter can be expressed as a simple equation which depends only on the observed distribution of baryonic mass. It is a challenge to understand how this very fine-tuned coupling between mass and light comes about. ", "introduction": "The masses of stellar disks and the distribution of mass in dark matter halos pose a coupled problem. Rotation curves provide good measures of the mass enclosed by disks. But it has been difficult to disentangle how much of this mass is in the stellar disk, and how much is in the dark matter halo. Consequently, both the mass of the stellar disk and the distribution of the dark matter, $\\rho(r)$, have been unclear. There have long been suggestions of a close connection between mass and light in spiral galaxies. Perhaps the most obvious is the Tully-Fisher relation (Tully \\& Fisher 1977). Beyond this global scaling relation, there are indications of a local coupling between mass and light (e.g., Rubin \\etal\\ 1985; Bahcall \\& Casertano 1985; Persic \\& Salucci 1991). One manifestation of this is in the efficacy of maximum disk (e.g., van Albada \\& Sancisi 1986) in describing the inner parts of rotation curves. If one scales up the stellar contribution to the rotation curves of high surface brightness (HSB) spirals to the maximum allowed by the data, one often finds a good match in the details (the ``bumps and wiggles'') between the shape of the rotation curve and that predicted by the observed stellar mass (e.g., Kalnajs 1983; Sellwood 1999; Palunas \\& Williams 2000). This only works out to some radius where dark matter must be invoked, but does suggest that the preponderance of the mass at small radii is stellar. Beyond that, it is often possible to scale up the gas component to explain the remainder of the rotation curve (Hoekstra, van Albada, \\& Sancisi 2001). Moreover, while some dark matter may be needed to stabilize disks, detailed analyses of disk stability frequently require rather heavy disks in order to drive the observed bars and spiral features (e.g., Athanassoula, Bosma, \\& Papaioannou 1987; Debattista \\& Sellwood 1998, 2000; Weiner, Sellwood, \\& Williams 2001; Fuchs 2003a; Bissantz, Englmaier, \\& Gerhard 2003; Kranz, Slyz, \\& Rix 2003). While these lines of evidence favor nearly maximal disks in HSB spirals, there are contradictory indications as well. The most significant of these is the lack of surface brightness (or scale length) residuals in the Tully-Fisher relation (Sprayberry \\etal\\ 1995; Zwaan \\etal\\ 1995; Hoffman \\etal\\ 1996; Tully \\& Verheijen 1997). The apparent lack of influence of the distribution of disk mass on the Tully-Fisher relation suggests that disks are submaximal (McGaugh \\& de Blok 1998a; Courteau \\& Rix 1999). However, galaxies which occupy the same location in the Tully-Fisher plane can have very different rotation curve shapes (de Blok \\& McGaugh 1996; Tully \\& Verheijen 1997). This excludes the simple hypothesis that galaxies of equal luminosity reside in identical halos with no significant influence from the disk. Recent data for low surface brightness (LSB) disks complicate matters further. These objects show large mass discrepancies down to small radii, implying that they are dark matter dominated with very submaximal disks (de Blok \\& McGaugh 1997). However, it is often formally possible to obtain a fit with something like a traditional maximum disk (where the peak velocity of the disk component is comparable to $V_{flat}$), albeit at the cost of absurdly high ($> 10\\; \\Msun/\\Lsun$) mass-to-light ratios (de Blok \\& McGaugh 1997; Swaters, Madore, \\& Trewhella 2000; McGaugh, Rubin, \\& de Blok 2001). As anticipated by McGaugh \\& de Blok (1998b), density wave analyses imply nearly maximal disks for LSB galaxies (Fuchs 2002, 2003b). These high mass-to-light ratios are unlikely for stellar populations, so one might consider a disk component of dark matter in addition to the usual halo. This seems contrived, and also causes problems with the baryonic Tully-Fisher relation (McGaugh \\etal\\ 2000; Bell \\& de Jong 2001). This relation between mass and rotation velocity works best, in the sense of having minimal scatter, for disk masses which are consistent with stellar population mass-to-light ratios (\\S 3). Maximal disks in LSB galaxies increase the scatter in the baryonic Tully-Fisher relation. Among these apparently contradictory lines of evidence, there is nevertheless a clear theme. The luminous and dark components are intimately linked. This is true not only in a global sense, but also in a local one. This might be paraphrased as Renzo's Rule: {\\it ``For any feature in the luminosity profile there is a corresponding feature in the rotation curve''} (Sancisi 2003). The distribution of baryonic mass is completely predictive of the distribution of dark matter, even in dark matter dominated LSB galaxies. Renzo's rule is an empirical statement which is mathematically encapsulated by MOND (Milgrom 1983). MOND is a modified force law hypothesized as an alternative to dark matter, and remains a viable possibility (Sanders \\& McGaugh 2002). Even if dark matter is correct (as is widely presumed), MOND is still useful as a compact description of the mass discrepancy in spirals (Sanders \\& Begeman 1994). In this paper, I show that the stellar mass-to-light ratios determined from MOND fits to rotation curves are optimal in a purely Newtonian sense. I derive a simple expression for the corresponding dark matter distribution, and generalize this to apply for {\\it any\\/} choice of stellar mass-to-light ratio. This expression provides a simple yet stringent test for dark matter theories which seek to explain rotation curves. ", "conclusions": "There is a strong relation between the distribution of baryonic and total mass in disk galaxies. The shape of the rotation curve predicted by the observed distribution of baryons is homologous to the observed rotation. The mass discrepancy, defined as the ratio of the gradients of the total to baryonic gravitational potential, can be described by a simple function of centripetal acceleration: \\begin{displaymath} \\Phi'/\\Phi_b' = {\\cal D}(x), \\end{displaymath} where $x = a/a_0$ and ${\\cal D}(x)$ is the inverse of equation~6. For $a > a_0 = 1.2 \\times 10^{-10}\\;{\\rm m}\\,{\\rm s}^{-2}$, there is no apparent need for dark matter. For $a < a_0$, the amount of dark matter increases systematically as acceleration declines. The empirical organization apparent in the data provides constraints on mass models which have not been considered in traditional disk-halo modeling. Making full use of the information present in the data provides a method for estimating the stellar mass-to-light ratio of spiral disks. These mass-to-light ratios are optimal in that they \\begin{itemize} \\item minimize the scatter in the mass discrepancy-acceleration diagram (Fig.~5), \\item minimize the scatter in the baryonic Tully-Fisher relation (Fig.~6), and \\item are consistent with the expectations of stellar population models (Fig.~7). \\end{itemize} Indeed, the mass-to-light ratios are consistent not only with the mean value anticipated for stellar populations, but they also reflect the expected trends with color and the band-pass dependent scatter about the mean color-\\ML\\ relation. It is hard to imagine that all this could be the case unless these mass-to-light ratios are essentially correct. This appears to solve the long standing problem of the absolute stellar mass of galactic disks. Once the stellar mass of a disk is fixed, the distribution of dark matter follows. This can be expressed as a simple function of observable quantities: \\begin{displaymath} \\Vh^2(r) = \\left[\\Q^{-1} {\\cal D}(x) -1\\right] \\Vst^2(r) + \\left[{\\cal D}(x) - 1\\right] \\Vg^2(r) \\end{displaymath} (equation~22). The dark matter distribution is completely specified by the observed baryonic matter distribution. This presents a serious fine-tuning problem for any theory of galaxy formation. The only freedom available in the detailed distribution of dark matter inferred from observation is encapsulated in a single parameter \\Q. This parameter allows for the unlikely possibility that the actual mass-to-light ratios of stars differs from the optimum value described here. With this limited freedom, the relation derived between baryonic and total mass specifies the dark matter distribution with complete generality. Explaining how this strong coupling between mass and light arises is a major challenge for galaxy formation theory. All we have done is restate, in a generalized Newtonian fashion, the well-established result that MOND fits rotation curves. The question, of course, is what this means. There are two independent issues here which are often mistakenly conflated. One is the unconventional theory known as MOND. The second is the empirical regularity of the data. If the effective force law apparent in the data is not in fact MOND, then we need to understand how it comes about in the context of dark matter. The possibilities are limited. Either \\begin{itemize} \\item MOND is essentially correct, or \\item Dark matter results in MOND-like behavior in disk galaxies. \\end{itemize} This is the same conclusion reached by Sanders \\& Begeman (1994) and McGaugh \\& de Blok (1998b). The improvement of the data since that time only makes this result more clear. If dark matter is correct, the tightness of the observed coupling between baryons and dark matter implies a very strong regulatory mechanism. Within the context of dark matter, there seem to be two basic options. Either \\begin{itemize} \\item the processes of galaxy formation lead to the observed coupling, or \\item there is a direct interaction between dark matter and baryons. \\end{itemize} In the first case, some combination of mundane astrophysical effects (e.g., adiabatic contraction, mergers, feedback) are presumed to be the root cause of the coupling. However, no clear mechanism is known which has the required effect. Indeed, it seems extremely unlikely that the chaotic processes of galaxy formation can give such a highly ordered result, much less the finely-tuned coupling between mass and light. Alternatively, the strong coupling between dynamical and baryonic mass might provide a hint about the nature of the dark matter itself. Rather than interacting with baryons only through gravity, there may be some direct interaction which results in the observed coupling. This amounts to a modification of the nature of the dark matter itself. One obviously wishes to retain the successes of CDM on large scales, but modify it to have the appropriate effect in individual galaxies. Many modifications of dark matter have been discussed in recent years (e.g., warm dark matter, self-interacting dark matter), motivated at least in part by the difficulties posed by rotation curves. Those proposed modifications which ignore the baryons would not seem to help. It is not adequate to change the dark matter properties in order to insert a soft core in dark matter halos; one must explain equation~22. A mechanism which provides for the direct connection observed between dark matter and baryons is unknown. Ideas along this line are only now being discussed (e.g., Piazza \\& Marinoni 2003), so it is too soon to judge whether they are viable. Irrespective of which possibility might seem to hold the most promise, there is clearly some important physics at work which has yet to be understood." }, "0403/astro-ph0403426_arXiv.txt": { "abstract": "{The eclipsing supersoft X-ray binary CAL 87 has been observed with Chandra on August 13/14, 2001 for nearly 100 ksec, covering two full orbital cycles and three eclipses. The shape of the eclipse light curve derived from the zeroth-order photons indicates that the size of the X-ray emission region is about 1.5 \\rsun. The ACIS/LETG spectrum is completely dominated by emission lines without any noticeable continuum. The brightest emission lines are significantly redshifted and double-peaked, suggestive of emanating in a 2000 km/s wind. We model the X-ray spectrum by a mixture of recombination and resonant scattering. This allows us to deduce the temperature and luminosity of the ionizing source to be $kT \\sim 50-100$ eV and $L_X \\sim 5 \\times 10^{37}$ erg/s. } \\addkeyword{binaries: close} \\addkeyword{Stars: individual: CAL 87} \\addkeyword{X-ray: stars} \\begin{document} ", "introduction": "\\label{sec:intro} CAL 87 was detected with the IPC onboard the Einstein X-ray observatory during the Columbia Astrophysical Laboratory survey of LMC (Long \\etal\\ 1981), and it was optically identified with an eclipsing binary in the Large Magellanic Cloud with an orbital period of 10.6 hrs (Pakull \\etal\\ 1988, Callanan \\etal\\ 1989, Cowley \\etal\\ 1990). A shallow X-ray eclipse was discovered with ROSAT (Schmidtke \\etal\\ 1993). Early attempts to model the optical light curve of CAL 87 by Callanan \\etal\\ (1989) already indicated an elevated accretion disk rim. This model has been extended by Schandl \\etal\\ (1997), and successfully describes the optical light curve of the CAL 87 binary system, composed of a primary (WD) with a mass of M$_1$ = 0.75 M$_{\\odot}$, placed at a distance of 2.2$\\times$10$^{11}$ cm from the mass-donating secondary star with a mass of M$_2$= 1.5 M$_{\\odot}$. The model of Schandl \\etal\\ (1997) includes (1) optical emission from the secondary star which is irradiated by the emission of the WD and (2) emission of an accretion disk with a thick rim and an optically thick, cold, clumpy spray produced by the high mass-flow rate of the accretion stream impinging on the disk (hot spot). This spray, moving around the disk, nicely reproduced the asymmetry in the optical light curve and the depth of the secondary dip. Based on the pre-Chandra, low-resolution X-ray spectra and the detection of emission up to 1 keV, CAL 87 has long been considered as one of the two hottest known SSS, thus spurring observations with all X-ray satellites since ROSAT (Parmar \\etal\\ 1997, Asai \\etal\\ 1998, Dotani \\etal\\ 2000, Ebisawa \\etal\\ 2001). We have exploited the unique spectral capability of the Chandra low-energy transmission grating (LETG) to study in detail the X-ray emission of the canonical supersoft source CAL 87. ", "conclusions": "The shape of the eclipse light curve, the lack of any continuum emission and the dominance of emission lines due to resonant scattering and recombination suggests that CAL 87's X-ray emission comes from an extended wind (or outflow), similar to the accretion-disk-corona low-mass X-ray binary 4U 1822-37, as already suggested by Schmidtke (1993). However, the plasma in 4U 1822-37 (i) has a much higher ionization as evidenced by the H- and He-like ions of Ne, Mg, Si, S and Fe which are completely absent in CAL 87 and (ii) is much more dominated by recombination, indicative of larger column densities than those in CAL 87. \\medskip \\outputfulladdresses The wind is probably not spherical, but rather has a conical shape with an opening angle of 120\\degs, extending on both sides of the accretion disk. During eclipse, the donor occults the wind on the southern disk side. The primary X-ray emission region in CAL 87 certainly has a temperature substantially lower than $\\sim$1 keV due to the lack of Fe L, Si XIII or Ne IX/Ne X lines. This rules out a neutron star or black hole (except for retrograde rotation) accretor, since they would exhibit a maximum temperature of the inner part of the accretion disk of 0.9--1.2 keV. Also, the outflow velocity of $\\sim$2000 km/s is consistent with a white dwarf primary. A corona with an ionization parameter and electron density ranging from $>$3000/10$^{12}$ cm$^{-3}$ near the accretor to $\\sim$30/10$^{11}$ cm$^{-3}$ near the outer boundary would scatter about 1\\% of the irradiated X-ray emission. This suggests that the intrinsic X-ray luminosity of CAL 87 is of order a few times 10$^{37}$ erg/s. Though the earlier SSS classification was based on a wrong interpretation of the X-ray spectrum, the new kT and L estimates from our LETG spectrum re-confirm the supersoft source nature of CAL 87." }, "0403/astro-ph0403389_arXiv.txt": { "abstract": "We present a survey of \\ion N5 and \\ion O6 (and where available \\ion C4) in the Galactic halo, using data from the {\\it Far Ultraviolet Spectroscopic Explorer} (FUSE) and the {\\it Hubble Space Telescope} (HST) along 34 sightlines. These ions are usually produced in nonequilibrium processes such as shocks, evaporative interfaces, or rapidly cooling gas, and thus trace the dynamics of the interstellar medium. Searching for global trends in integrated and velocity-resolved column density ratios, we find large variations in most measures, with some evidence for a systematic trend of higher ionization (lower \\ion N5/\\ion O6 column density ratio) at larger positive line-of-sight velocities. The slopes of log[N(\\N)/N(\\O)] per unit velocity range from $-0.015$ to $+0.005$, with a mean of $-0.0032 \\pm 0.0022$(r)$\\pm 0.0014$(sys) dex~(km~s\\ts{-1})\\ts{-1}. We compare this dataset with models of velocity-resolved high-ion signatures of several common physical structures. The dispersion of the ratios, \\ion O6/\\ion N5/\\ion C4, supports the growing belief that no single model can account for hot halo gas, and in fact some models predict much stronger trends than are observed. It is important to understand the signatures of different physical structures to interpret specific lines of sight and future global surveys. ", "introduction": "We are interested in understanding how the dynamics of the interstellar medium (ISM) of galaxies determines how the energy and matter released by stars are redistributed. In a previous paper \\citep{remy04}, hereafter denoted Paper~I, we investigated theoretical models of the ionization ratios of Li-like absorbers, \\ion N5, \\ion O6, \\ion C4, in the Galactic halo. These ions are believed to form in nonequilibrium in shocks, evaporative interfaces, or rapidly cooling gas, all of which trace the dynamics of the interstellar medium. As a useful new diagnostic, in that paper we focused on {\\it velocity-resolved} signatures of several common physical structures: (1) a cooling Galactic fountain flow that rises, cools, and recombines as it returns to the disk; (2) shocks moving toward the observer; (3) a conductive interface with the observer located in the hotter gas. In this paper, we present a survey of 34 sightlines observed through the Galactic halo by the {\\it Far Ultraviolet Spectroscopic Explorer} (FUSE) and the {\\it Hubble Space Telescope} (HST) in the Li-like absorption lines of \\ion O6, \\ion N5, and sometimes \\ion C4. We report the integrated column densities of these three ions and their ratios. We also focus, as in Paper~I, on the new, and potentially useful diagnostic provided by the velocity dependence of these ion ratios. The observable velocity-ionization trends are weak, because even strong trends are washed out by the large thermal width of the gas at different parts of the flow. Additional confusion results since the long sightlines almost definitely pass through multiple structures. We report an indication of a weak trend of decreasing N(\\N)/N(\\O) at more positive velocities. The distribution of this slope is broad, with a mean value, $-0.0032\\pm0.0022$(r) $\\pm0.0014$(sys)~dex~(km~s\\ts{-1})\\ts{-1}, displaced somewhat from zero. In this paper, \\S~\\ref{data} describes the data calibration and analysis. The basic observational results are described in \\S~\\ref{results}. Section \\ref{interpretation} describes models of the dynamical signatures of the Li-like ions, and interpretation. ", "conclusions": "We consider the diagnostic power of velocity-resolved column density ratios in understanding the Galactic halo. Column density ratios of Li-like ions in the Galaxy are useful to diagnose the physical formation mechanism of the gas and to study the interstellar gas cycle, and a survey of these ions can reveal general trends. We find, in a survey of sightlines observed with FUSE and HST, that the distribution of N(\\N) and N(\\O) in the halo does not appear to favor a dominant physical production mechanism. We find a possible weak trend of decreasing N(\\N)/N(\\O) at more positive velocities ($-$0.0032$\\pm$0.0022(r)$\\pm$0.0014(sys)~dex~(km~s\\ts{-1})\\ts{-1}, Figure \\ref{histo-slope}). The weakness of this trend also argues against halo sightlines dominated by a single structure, e.g. the Local Bubble interface. In Paper~I we presented models of interfaces and cooling nonequilibrium gas, and the velocity-resolved N(\\N)/N(\\O) signatures of each. Here, we consider the model of smoothly distributed gas corotating on cylinders with the Galactic disk, and implications for the scale heights of \\N\\ and \\O. Observable velocity-ionization trends are weak, because even very strong trends are washed out by the large thermal width of the gas at different parts of the flow. Additional confusion results because the long sightlines almost definitely pass through multiple structures. In fact, the dispersion of N(\\N)/N(\\O), both integrated and velocity-resolved, clearly indicates that no single production scenario known to date can completely explain the Galactic halo. To truly understand the physical production of Li-like ions in the halo, one needs to analyze gas in localized areas of physical space, rather than velocity space. Absorption spectroscopy towards many halo stars with close angular separation and different distances could help to isolate gas at a specific altitude. Similarly, the gas above known superbubble shells or chimneys could be isolated. These observations have a greater chance of distinguishing between models of hot gas production than observations along long lines of sight." }, "0403/astro-ph0403176_arXiv.txt": { "abstract": "We have studied published data from the {\\it Yohkoh} solar X--ray mission, with the purpose of searching for signals from radiative decays of new, as yet undiscovered massive neutral particles. This search is based on the prediction that solar axions of the Kaluza--Klein type should result in the emission of X--rays from the Sun direction beyond the limb with a characteristic radial distribution. These X--rays should be observed more easily during periods of quiet Sun. An additional signature is the observed emission of hard X--rays by SMM, NEAR and RHESSI. The recent observation made by RHESSI of a continuous emission from the non--flaring Sun of X--rays in the 3 to $\\sim 15$~keV range fits the generic axion scenario. This work also suggests new analyses of existing data, in order to exclude instrumental effects; it provides the rationale for targeted observations with present and upcoming (solar) X--ray telescopes, which can provide the final answer on the nature of the signals considered here. Such measurements become more promising during the forthcoming solar cycle minimum with an increased number of quiet Sun periods. ", "introduction": "In order to solve the strong CP problem (why the strong interaction contrary to weak interactions does not violate CP symmetry), a new neutral particle, the axion, with spin--parity $0^-$ has been invented \\citep[for recent review articles see, e.g.][]{raffeltg99,raffeltg00,bradley}. Axions, along with Weakly Interacting Massive Particles (WIMPs), are the two leading particle candidates for dark matter in the Universe. If axions exist, they should be abundantly produced inside the solar core. An axion can be seen as a very light neutral pion ($\\pi ^0$), with restmass $\\sim 10^{-3\\pm 3}$ eV/c$^2$. However, while the $\\pi ^0$ interacts strongly and decays to two photons with a mean lifetime of $\\sim 10^{-16}$ s, the axion interacts very feebly with matter and is expected to decay to two photons ($a\\rightarrow 2\\gamma$) with a lifetime much longer than the age of the Universe. Therefore, detection techniques utilise the axion interaction with the electric field of atoms of underground detectors or strong magnetic fields, in order to (coherently) convert them to real photons. In other words, electric or magnetic fields play the role of a catalyser, which can transform axions into detectable photons. If very light axions ($m \\leq $ few eV/$c^2$) are produced inside the Sun, their thermal energy peaks at $\\sim 4.2$ keV and they are ultra-relativistic. They mostly stream out of the Sun and can have an impact only on the evolution of a Star through the additional escaping energy (as it happens with neutrinos). However, in recent theories of extra--dimensions, proposed as extensions of the Standard Model, the `conventional', almost massless axions become as massive as the reaction energies involved. In the case of the solar axions, the expected mass spectrum of all the excited Kaluza--Klein (KK) states extends all the way to $\\sim 10$ keV/c$^2$ \\citep{kk,dl,dilella}. These high KK--masses imply a relatively shorter lifetime ($\\tau \\sim 10^{20}$s), because of the $\\tau \\sim m^{-3}$ dependence. The underlying axion--photon--photon coupling constant, $g_{a\\gamma \\gamma}$, which defines the interaction cross-section with ordinary matter, is the same for the `conventional' almost massless axion and for the massive KK--axions. In this work, the KK--axions are taken as a generic example of particles which can be created inside the hot solar core. A small fraction of them ($\\sim 10^{-7}$, as it has been estimated by \\citet[hereafter DZ03]{dilella}), are extremely nonrelativistic and they can be gravitationally trapped by the Sun itself in orbits where they accumulate over cosmic times. As shown in DZ03, their density increases enormously near the solar surface. The estimated mean distance of the KK--axion population from the Sun surface is $\\sim 6.2\\,R_{\\odot}$. If axions of the KK--type are gravitationally trapped and decay to two photons,% \\footnote{Following the decay mode $a\\rightarrow \\gamma \\gamma$, the two photons have the same energy and are emitted back--to--back, because of momentum conservation, since they are highly non-relativistic.} then the observed X--ray surface brightness from the solar disk and limb can be a signature of the solar axion scenario. In this work, we discuss the expected surface density profile of the derived axion halo around the Sun (DZ03), and we compare it with published data taken by the {\\it Yohkoh} soft X--ray telescope from the diffuse emission of two quiet Sun observations. In addition, the measured continuous emission of hard X--rays (below $\\sim15\\mbox{ keV}$) from the non--flaring Sun during the past $\\sim25$ years by different orbiting X--ray detectors is also considered. ", "conclusions": "Our work suggests that the importance of quiet--Sun solar X--rays should not be overlooked. This paper should provide motivation to the solar X--ray community to follow experimentally the arguments presented here and to reduce the uncertainties of relevant observations. The period during the (next) solar cycle minimum should provide still cleaner quiet Sun conditions. For example, offpointing observations with the RHESSI solar X-ray telescope can result to interesting data above $\\sim 3$ keV.~Even the pointing tests of the spacecraft on the 8 and 23 May 2003 might well contain useful solar data, in spite of the announcement that `data in this period should not be used for solar studies'\\,% \\footnote{http://hesperia.gsfc.nasa.gov/hessi/ and http://hessi.ssl.berkeley.edu/$\\sim$dsmith/hessi/HME.html} $\\!\\!$. Similarly, we suggest that the announced offpointing to the Crab Nebula from 16 June 2003 should be evaluated in the light of this work. From Figure 2 and the energy threshold of the RHESSI detectors, it follows that the relevant angular extension from the surface of the Sun is approximately $\\theta \\leq 2^{\\circ}$. Other non--solar X--ray telescopes can be of potential use for this work, provided they can reduce the elongation angle with the Sun. Then, instead of searching for solar axions only here at the site of the Earth, (solar) X--ray telescopes could provide new physics results associated with the solar X--ray halo. At present, these orbiting instruments are highly sensitive to X--rays below $\\sim 10$ keV, which is within the region-of-interest for many astrophysical objects. In order to explain the experimental approach of this work in a model independent way, an exotic effect like the axion scenario of this paper can be searched for in the following type of solar observations, provided conventional solar dynamical effects can be excluded: a) an off--pointing observation results to an increased radiation level while approaching the Sun. This should be the manifestation of a halo around the Sun {\\it whatever} its radial distribution, and/or b) the measured level of {\\it any} cosmic radiation from the blank sky does not diminish completely (according to the performance of the telescope) while pointing towards the solar disk, i.e.\\ solar shadow effect is missing. The search for such a residual radiation from the solar disk direction might be the most sensitive approach, especially at photon energies above a few keV, where practically no emission is expected from the quiet Sun. {\\it In summary}, the radial distributions from the re--considered two {\\it Yohkoh} X--ray observations of the quiet Sun (including the derived inward heat flux in the solar atmosphere of {\\it some} nonthermal energy deposition beyond $\\sim 1$ solar radius from the Sun surface) can be reconciled with a halo of decaying massive particles near the solar surface. New analyses of existing data may definitely clarify that instrumental scattering effects from bright points on the solar disk are small. More extended radial X--ray distributions (preferentially during a solar cycle minimum) can provide important information for the quiet Sun. An additional and independent evidence in favour of the axion scenario is the observed {\\it continuous} emission of X-rays from the non-flaring Sun in the 3 to $\\sim 15$ keV range. All these observations, when considered together, might suggest a non--conventional mechanism of the type assumed here. New X--ray measurements could clarify the nature of these relatively high energy effects around the Sun, for which an alternative conventional explanation is still missing." }, "0403/astro-ph0403495_arXiv.txt": { "abstract": "High resolution spectral studies were undertaken at orbital phases ($\\phi$) 0, 0.25 and 0.5 on the high-mass X-ray binary (HMXB) Vela X-1 using archival Chandra data. We present (a) the first detailed analysis of the multiple strong narrow emission lines present in $\\phi$ 0.5 (b) an analysis of the absorption of the continuum in $\\phi$ 0.5, and (c) the first detection of narrow emission and absorption lines in $\\phi$ 0.25. Multiple fluorescent and H-and He-like emission lines in the band 1.6 - 20 Angstrom (\\AA) in eclipse are partially obscured at $\\phi$ 0.25 by the X-ray continuum. The $\\phi$ 0.25 spectrum displays 3 triplets, 2 with a blue-shifted resonance (r) line in absorption and the intercombination (i) and forbidden (f) lines in emission, and shows in absorption other blue-shifted lines seen in emission in eclipse. At $\\phi$ 0.5 the soft X-ray continuum diminishes revealing an \"eclipse-like\" spectrum, however line flux values are around 13-fold those in eclipse. We conclude the narrow emission lines in Vela X-1 become apparent when the continuum is blocked from line of sight, either by eclipse or by scattering and/or absorption from a wake or cloud. The H-and He-like lines arise in warm photoionised regions in the stellar wind, while the fluorescent lines (including a Ni K$\\alpha$ line) are produced in cooler clumps of gas outside these regions. Absorption of the 5-13 $\\mbox{\\AA} $ continuum at $\\phi$ 0.5 may be caused by an accretion wake comprised of dense stagnant photoionized plasma inside a Stromgren zone. Multiple fluorescent emission lines may be a common feature of the supergiant category of HMXBs. ", "introduction": "High-mass X-ray binaries (HMXBs) including Vela X-1 have served to illuminate the evolution of giant binary stars, to probe their radiation-driven winds, and to study accretion processes and the fundamental parameters of compact objects. The 130 known HMXBs fall into 2 categories, Be/X-ray and supergiant systems \\citep{liu00}. Vela X-1 is one of 11 supergiant systems (which include GX 301-2 and 4U 1700-37), while the vast majority (around 80\\%) of HMXBs are Be/X-ray systems detected as transients \\citep{kap01,whi95}. Supergiant systems have a relatively short life of 10,000 years before the secondary star explodes in a supernova leaving a neutron star binary or 2 single neutron stars \\citep{kap01}. Once a massive main-sequence star evolves to a supergiant the Roche lobe may become filled and accretion flow becomes sufficient to power a strong X-ray source. Vela X-1 is an eclipsing system at an estimated distance of 1.9 kpc \\citep{sad85} containing an B0.5Ib supergiant (mass 23 $M_\\odot$ and radius 34 $R_\\odot$) and a wind-fed pulsar with a period of 283 s \\citep{nag89}. The neutron star (NS) orbits in a period of 8.96 days very close to the non-synchronously rotating surface of the supergiant with a semi-major axis 53.6 $R_\\odot$\\citep{bar01}, so it is fully embedded in the acceleration zone of the stellar wind. The NS mass of 1.9 $M_\\odot$ is the highest known \\citep{bar01,van95} and the system has been studied at infrared \\citep{hut02}, UV \\citep{sad85, vanl01}, gamma \\citep{rau94} and at optical \\citep{van95, kap94} wavelengths. It has not been detected at radio frequency (detection limit at mean cm radio flux density (mJy) $<0.2$ \\citep{fen00}. Orbital modulation of the spectra have been noted in the UV, X-ray, and optical wavelengths. \\citet{sad85} noted absorption of \\ion{Al}{3} and \\ion{Fe}{3} in the UV, limited to the orbital phase ($\\phi$) 0.5 and higher, and postulated the presence of a trailing wake. \\citet{bes75} reported an absorption component in the H$\\alpha$ spectrum around $\\phi$ 0.6. \\citet{kap94} postulated that an absorption component in the optical spectrum resulted from a photoionization wake that trails the X-ray source, rather than the accretion wake which does not sufficiently obscure the supergiant. \\citet{smi01} studied an asymmetric dip in the UV lightcurve at $\\phi$ 0.46-0.7 and concluded spectral ratios (predip/dip) contain information about the mechanism responsible for the variability. Absorption of soft X-rays during the $\\phi$ 0.5 and higher was first reported in Tenma data \\citep{nag89}. \\citet{fel96} showed that a trailing photoionization wake can explain the asymmetry in the X-ray band during the eclipse ingress and egress. \\citet{sak99} using ASCA data demonstrated that the spectrum in eclipse was comprised of recombination lines and radiative combination continua (RRC) produced by photoionization in the stellar wind; they suggested the fluorescent K-shell lines from near-neutral atoms indicated that the X-ray -irradiated portion of the wind in Vela X-1 consists of cool dense clumps in a hotter, more ionized gas. \\citet{shu02a} presented the first high-resolution spectrum during eclipse, and using RRC (including \\ion{Ne}{10} RRC) deduced a temperature of 1.2 $\\times 10^5$ K for the plasma. The spectrum showed clear evidence of photoionization processes, and the resonance (r) lines in the He-like triplets were of roughly equal strength to the forbidden (f) lines, consistent with resonant scattering. We unexpectedly found that in $\\phi$ 0.5 the emission lines were stronger than in eclipse by an order of magnitude, associated with the near disappearance of the soft X-ray continuum. This result, together with the unusual finding of multiple fluorescent emission lines, led us to further investigate emission line production in HMXBs by a study of Vela X-1 comparing eclipse, $\\phi$ 0.25 and $\\phi$ 0.5 using high-resolution Chandra datasets. We outline the data reduction proceedure in section 2, give our results in section 3 and present the discussion in section 4. ", "conclusions": "X-ray emission in Vela X-1 results from capture and accretion of gas in the stellar wind of the supergiant star onto the hot spot on the NS. Wind material departs uniformly from the surface of the star, and accretes onto the NS if it passes within a critical distance forming an accretion column. Adjacent material that is perturbed in its flow but is not accreted forms an accretion wake. The X-rays ionize and heat the surrounding gas, and are reprocessed in the stellar wind. Spectra at orbital phases 0, 0.25 and 0.5 reflect the processes of mass loss in the stellar wind of the supergiant and mass accretion onto the NS. The Doppler shifts at $\\phi$ 0.5 indicate slowing of the supergiant's wind velocity in the vicinity of the NS, consistent with X-ray photoionization and destruction of ions with strong UV resonance transitions that otherwise drive the stellar outflow. The X-ray source creates a Stromgren zone in which the high photoionization conditions effectively quench the driving force and acceleration of the wind. Spectra at $\\phi$ 0.25 reveal simultaneous evidence of emission lines produced by scattering, and absorption lines that are produced in different regions of the binary system. Narrow emission lines in Vela X-1 become apparent when the continuum is blocked from line of sight, either by eclipse or by scattering and/or absorption from a wake or cloud. The X-ray source is very compact and is readily occluded by the accretion wake at $\\phi$ 0.5, accounting for the absorption of soft continuum. Absorption of the continuum at $\\phi$ 0.5 has allowed detection of scattered radiation and emission lines whose strength is an order of magnitude higher than those in the eclipse phase. The eclipse and 0.5 spectra have a similiar appearance, and the ratios of lines in the He-like triplets are consistent with photoionization and resonant scattering (involving photoexcitation) in a PI plasma. Multiple fluorescent emission lines have been detected in 3 of 5 supergiant HMXBs for which high resolution X-ray spectra are available. This study suggests fluorescent emission from externally X-ray illuminated matter may occur from elements other than neutral iron if the direct radiation is blocked and only reflected radiation is visible, when neutral ions of Ne, Mg, Si, S, Ar, Ca, Cr and Ni may be detected. Differences in Doppler shifts of various lines indicate different origins of lines, from direct radiation from the surface of the NS (lines seen in $\\phi$ 0.25 in absorption) and from adjacent regions in the stellar wind from scattered radiation." }, "0403/astro-ph0403206_arXiv.txt": { "abstract": "We investigate in detail the mass distribution obtained by means of high resolution rotation curves of 25 galaxies of different morphological types. The dark matter contribution to the circular rotation velocity is well-described by resorting to a dark component whose density shows an inner core, i.e. a central constant density region. We find a very strong correlation between the core radius size $R_C$ and the stellar exponential scale length $R_D$: $R_C \\simeq 13 \\ (\\frac {R_D} {5\\ {\\rm kpc}})^{1.05} \\ {\\rm kpc} $, and between $R_C$ and the galaxy dynamical mass at this distance, $M_{dyn}(R_C)$. These relationships would not be expected if the core radii were the product of an incorrect decomposition procedure, or the biased result of wrong or misunderstood observational data. The very strong correlation between the dark and luminous scale lengths found here seems to hold also for different Hubble types and opens new scenarios for the nature of the dark matter in galaxies. ", "introduction": "\\label{sec:introduction} In spite of a large amount of knowledge gathered over the past 20 years on the phenomenon of Dark Matter (DM) in Galaxies, it is only recently that the attention has been focused on its radial density profile. Recall that, according to high-resolution N-body simulations in the $\\Lambda$-CDM framework, the dark halo density, characterized by the virial halo mass as free parameter, is described by an inner power--law cusp (but see Ricotti 2003): \\begin{equation} \\rho_{\\rm NFW}(r) = \\frac{\\rho_s}{(r/r_s)(1+r/r_s)^2} \\label{eq:rho_nfw} \\end{equation} where $r_s$ and $\\rho_s$ are related parameters (Navarro, Frenk and White, 1996b, hereafter NFW). The DM contribution to the circular velocity can be written as: \\begin{equation} V_{\\rm NFW}^2(r)= V_{vir}^2 \\frac{c}{A(c)} \\frac {A(x)}{x} \\end{equation} where $x \\equiv c r/R_{vir}$, $c \\equiv R_{vir}/r_s$ is the concentration parameter, and $A(x)\\equiv \\ln (1+x) - x/(1+x)$. We also have that $V^2_{vir}= G M_{vir}/ R_{vir}$, $\\rho_s \\simeq 101/3 \\ \\rho_{crit} \\ c^3 \\ / (\\ln(1+c)-c/(1+c))$ and $M_{vir} = \\frac{4 \\pi}{3} \\delta_{vir} \\rho_{mean} R^3_{vir}$. \\footnote {$\\rho_{crit}$ is the critical density of the universe, $\\delta_{vir} \\simeq 337$ is the virial overdensity and $\\rho_{mean}$ is the mean universal density at the galaxy's redshift. The values for the numerical constants are valid for a $\\Lambda$-CDM cosmogony with $\\Omega_0$=0.3 and $\\Omega_\\Lambda$=0.7.} A number of reliable mass modellings, obtained for spirals and low surface brightness galaxies (LSBs) has supported the early claim (Moore, 1994) that dark halos around disk galaxies have a central density distribution much shallower than the NFW one (e.g. Gentile et al., 2004 and references therein; Weldrake, de Blok \\& Walter, 2003; de Blok \\& Bosma, 2002). The first studies on elliptical galaxies seem to indicate that also these objects share the same phenomenon (Gerhard et al., 2001; Borriello et al., 2003 ). This ``galaxy by galaxy'' comparison between the predicted $\\Lambda$-CDM density distribution and those actually detected for the dark halos around disk galaxies has been the main goal of several published works. However, although the study of the discrepancy between data and theoretical predictions remains necessary, we believe that time has come that we investigate {\\it per se} the distribution of DM around galaxies, independently of the existing cosmological implications. In fact, the hot debate on possible falsification of $\\Lambda$-CDM on galactic scales has carried the research off from the original topic of understanding the {\\it DM phenomenon} in virialized systems. Let us stress that a {\\it direct} knowledge on the presence, nature and interaction with baryons of the dark galactic component is still very rough and limited, unlike the complex and refined scenario that theory and simulations have put forward. In the decomposition of rotation curves (RC's), the circular velocity $V(r)$ is best-fit modelled as the sum of a baryonic component that includes a stellar and a gaseous disk, and a spherical dark halo: \\begin{equation} V^2(r)=V_D^2(r)+V^2_{gas}(r)+V^2_H(r) \\end{equation} where the labels D, gas and H refers to the corresponding components. It is common to represent the dark halo contribution to the circular velocity by means of the 2-parameter function describing the circular velocity of a pseudo-isothermal (PI) halo: \\begin{equation} V^2_{H,PI}(r) = 4\\pi G \\rho_0 R_C^2 \\left( 1 - \\frac{R_{C}}{r} {\\rm arctan} \\frac{r}{R_{C}} \\right). \\label{eq:pi} \\end{equation} The central density $\\rho_0$ and the halo {\\it density core} radius $R_C$ are free parameters to be tuned to fit $V^2(r)$. Persic, Salucci \\& Stel (1996, PSS hereafter) introduced, for the dark halo velocity contribution, an equally simple function, \\begin{equation} V^2_{H, URC}(r)=(1-\\beta) V_{opt}^2 r^2 (1+a^2)/(r^2+a^2 R_{opt}^2) \\label{eq:urc} \\end{equation} to fit, once added to the disk contribution, the Universal Rotation Curve (URC) of Spirals, i.e. the ensemble of synthetic RC's, each one derived from a large number of individual RC's belonging to objects with a luminosity falling inside a fixed range (see Section 4 of PSS for details). The two free parameters are: the {\\it velocity} core radius $a R_{opt}$, and $(1-\\beta) \\ V^2_{opt}$, the halo velocity amplitude at $R_{opt}$.\\footnote{$R_{opt}$ is the radius encompassing 83\\% of the total luminosity of the galaxy. In the case of a (stellar) exponential thin disk $R_{opt}$ is 3.2 times the disk scale length $R_D$.} For the aim of this work, in the region of interest, $1/3 R_{opt} \\leq r \\leq R_{opt}$, the PI circular velocity is indistinguishable from the {\\it halo } URC of Eq. 5. In fact, the transformation law is obtained from the known quantities $R_{opt}$ and $V_{opt}$, and by setting the PI free parameters $R_C$ and $\\rho_0$ in Eq. 4 at the values $R_C= a\\, R_{opt}/1.45$ and $\\rho_0= [ 4\\pi G R_C^2 ( 1 - \\frac{R_{C}}{R_{opt}} {\\rm arctan} \\frac{R_{opt}}{R_{C}})]^{-1} (1-\\beta) V_{opt}^2$. Let us notice that $ V_{H, PI}$ and $V_{H,URC}$ agree with each other within few percent at most, a discrepancy that is completely negligible, in view of the large radial variations of $V(r)$ inside galaxies and of the large differences among the halo velocity profiles predicted by different mass models. Moreover, the velocity profiles of Eqs. \\ref{eq:pi} and \\ref{eq:urc}, once their free parameters have been properly set, can mimic in the regions where data are available a number of ``theoretical'' rotation curves, not all necessarily consistent with observations.\\footnote{e.g. the NFW circular velocity with $r_s=R_{opt}$, in the radial range, $0.3< r/r_s<1.1$ is reproduced, within a negligible few percent discrepancy, by $V_{H, URC} (r; a=0.33 )$ and/or by $V_{H, PI}(r; R_C=0.2 R_{opt})$.} Therefore, with the obvious caveat that suitable data are available, by investigating the properties of the density distribution of the dark halos around galaxies by means of the PI/URC profile we neither assume nor reject the theoretical $\\Lambda$-CDM scenario. On the contrary, by adopting certain values for the free parameters of the chosen halo density profile in order to match the observations, we will be able to probe a density cusp, or a core or even something in between. In this work, we first obtain the density profile of DM halos around a significant number of galaxies of different Hubble types with high-quality kinematic data. The intrinsic high quality of the data will allow us to disentangle the available kinematics in its dark and luminous contributions and to constrain the DM halo parameters allowing us to investigate the existence of a link between their structural properties and those of the ordinary baryonic matter (i.e. stars and gas). The sample of data used in our analysis is presented and discussed in Sect. \\ref{sec:sample}, while in Sect. \\ref{sec:model} we describe the mass modelling. In Sect. \\ref{sec:results} we will look for a relationship between the dark and stellar structural parameters. Finally the conclusions will be drawn in Sect. \\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} There is a mounting evidence that the kinematics of disk galaxies can be explained in terms of the standard disk + halo components {\\it only } if the density distribution of the latter decreases with radius very slowly from the center out to the radius $R_C$ inside which, instead, the density of CDM halos is predicted to fall as a power law with an exponent between -1 and -2. Here we have investigated this ``observational\" dark matter scale $R_C$ that is absent in the current theory of galaxy formation. We found that it correlates tightly with the exponential thin disk scale length $R_D$ and with the amount of gravitational mass that it encompasses $M_{dyn}(R_C)$. The high values for the correlation number, the smallness of the scatter around the relations and especially the fact that the pairs $R_C$ and $R_D$, as well as $R_C$ and $M_{dyn}(R_C)$, are measured/derived independently from each others, ensure that we are dealing with {\\it real and physical} relationships among physical quantities. Three different levels of consequences follow. First, the claims according to which core radii arise as a consequence of 1) serious observational errors, 2) peculiar or biased kinematics, 3) wrong mass modelling that, until today, have been { \\it by-passed} by considering a number of test-cases, can be now ruled out. This, on the basis that human/observational error hardly relate to intrinsic properties of galaxies. As an example the suggestion that observers have badly missed the galaxy center and doing so artificially created a core, cannot explain the additional evidence that this core is found to correlate with the galaxy disk scale length within the extraordinarily small scatter of 0.16 dex. The second level realises that the core radius, the quantity that defines the unexplained feature of the distribution of the dark matter, correlates with the main quantity controlling the distribution of the luminous matter as well as the total mass within this radius. This indicates that the DM density in galaxies has been shaped by a dark--to--luminous matter coupling, a very challenging task if the former is collisionless. However, let us recall that, at least in dwarf systems, according to Navarro et al. (1996a), large mass outflows from galaxies, arising from supernovae explosions, can modify an initial NFW profile into a cored one, for which $R_C \\propto (M_D/R_D)^{0.5}$. From the results of the mass modelling given in PSS, we find $M_D \\propto R_D^3$, in excellent agreement with relationship (9). In any case, our result points toward some new physics of which it will be a crucial benchmark. The third level is the realization that the core radius vs disk scale length relationship is Hubble type free, at least within a band of 0.16 dex. For the objects in our sample normal/dwarf spirals and LSB lay on the same $\\log R_C$ vs $\\log R_D$ relationship. Moreover, pioneering analysis on the DM distribution in ellipticals (Gerhard et al. 2001) finds halo core radii of sizes proportional to the half light radii (that correspond to 1.67 $R_D$): by analysing the mass model parameters of 21 Giant Ellipticals (shown in Fig 18 Gerhard et al. (2001), and made available by the authors) we find: \\begin{equation} \\log R_C= (1.1 \\pm 0.2 )\\ \\log R_D +(0.1 \\pm 0.4) \\end{equation} in very good agreement with Eq. (9). We caution that the sample of this work is quite limited to fully investigate the morphological dependence of DM properties and that in Gerhard et al.(2001) the data quality is barely sufficient for the intrinsically complicated mass modelling. Nevertheless, the emerging picture is truly impressive, especially when we consider that core radii of galaxies of different Hubble types do not correlate with galaxy luminosity: e.g. a similar value for $R_C$ is found in a $10^{11} L_{\\odot,r}$ spiral, in a $10^{10} L_{\\odot,r}$ LSB and in a $10^{12} L_{\\odot,r}$ giant elliptical. The core radius seems to uniquely relate to the exponential rate with which the stellar surface density decreases with radius. The next step of this line of research post-$\\Lambda$-CDM crisis would be to thoroughly investigate this DM distribution property in galaxies of different size and Hubble Types (Salucci et al., 2004) and the brightest galaxies because of the slightly non-linear relationship found between $R_C$ and the stellar scale length." }, "0403/astro-ph0403030_arXiv.txt": { "abstract": "We present the first results of the timing analysis of our {\\emph{RXTE/INTEGRAL}} monitoring campaign on GRS~1915+105. Over the 6 already performed {\\emph{RXTE}} observations, we study the presence of Low Frequency QPO (LFQPO), and their energetic dependences. In a view to understand the QPO phenomenon, we compare the QPO properties to the spectral behaviour of the source. We propose that part of the compact jet detected during multi-wavelength observations, could produce a significant amount of hard X-rays, and hence explain the energy dependence of the amplitude of the QPOs. ", "introduction": "\\begin{table*} \\begin{tabular}{|ccccc|} \\hline Obs. sequence \\# & Date (MJD)& Good times & Revolution \\# & {\\emph{INTEGRAL}} ref. \\\\ ({\\emph{RXTE}}) & & (s) & ({\\emph{INTEGRAL}}) & \\\\ \\hline 1 & 6-7 March 2003 (52704-05)&15768 & Rev. 48 & Hannikainen et al. 2003 \\\\ 2 & 2 April (52731)&9300 & Rev. 57 & Fuchs et al. 2003 \\\\ 3 & 9-10 April (52738-39)&25360 & Rev. 59 & Hannikainen et al. 2004 \\\\ 4 & 9 May (52768)& 14000& Rev. 69 & Hannikainen et al. 2004 \\\\ 5 & 2 November (52945) & & Rev. 122 & Solar Flares\\\\ 6 & 22-23 November(52965-66) &36100& Rev 135 & Hannikainen et al. 2004 \\\\ \\hline \\end{tabular} \\caption{log of the {\\emph{RXTE}} observation and contemporaneous {\\emph{INTEGRAL}} revolutions reported in this paper} \\label{tab:log} \\end{table*} The log of the {\\emph{RXTE}} observations analysed in this paper is reported in Table \\ref{tab:log}. All of them were simultaneous with {\\emph{INTEGRAL}} and other ground based observations. The first observation shows a new class of variability (Hannikainen et al. 2003), while during Obs. 2,3,4, the source has a steady flux in the X/Gamma rays. It shows a strong QPO, a powerful compact jet during Obs. 2 (Fuchs et al. 2003), and a high level of radio emission during the 2 other observations (Hannikainen et al. 2004, these proceedings). ", "conclusions": "We found that the radio loud hard state (a.k.a. Class $\\chi_1$-$\\chi_3$) does not lead to sensible parameters when fitted with a multi colour disc black body and a power-law (see Muno et al. 2001, Trudolyubov 2001) . A broken power-law represents the 3-300 keV spectra well. \\\\ \\indent\tDuring the 3 steady observations high level of radio emission is observed (see Fuchs et al. 2003, Hannikainen et al. these proceedings). The hard X-rays above 20 keV may originate from the jet (as expected see Markoff et al. 2003; Corbel et al. 2003). The spectra of GRS 1915+105, would then fit better in the standard picture of micro-quasars states.\\\\ \\indent\tOur analysis of the spectra of LFQPOs confirms the presence of a cut-off in their energy dependence (Tomsick \\& Kaaret 2001; Rodriguez et al. 2002), with an evolving energy from $\\sim$15-20 keV in Obs. 2 to a value $ >$25 keV in the remaining Obs. (Fig.\\ref{fig:amplitude}). Although the QPO cut-off is needed at some point (otherwise its amplitude would grow indefinitely), its evolution is unclear. It can be easily understood, however, if the jet contributes significantly to the hard X-ray, and its flux is not modulated on such a short time scale. \\\\ \\indent\tThe fact that a high level of radio emission (at 15 GHz) is found during obs.2, when the cut-off is clearly detected in the QPO spectrum, is compatible with this interpretation. We thus propose that a part of the hard X-ray is emitted by the compact jet, explaining both the spectral behaviour of the source and the QPO ``spectra''." }, "0403/astro-ph0403564.txt": { "abstract": "Motivated by observations of inner halo satellite remnants like the Sgr and \\omegacen, we develop fully analytical models to study the orbital decay and tidal massloss of satellites on eccentric orbits in an isothermal potential of a host galaxy halo. The orbital decay rate is often severely overestimated if applying the ChandraSekhar's formula without correcting for (a) the evaporation and tidal loss of the satellite and (b) the contraction of satellite orbits due to adiabatic growth of the host galaxy potential over the Hubble time. As a satellite migrates inwards, the increasing halo density affects the dynamical friction in two opposite ways: (1) it boosts the number of halo particles swept in the satellite's gravitational \"wake\", hence increasing the drag on the satellite, and (2) it boosts the tide which \"peels off\" the satellite, and reduces the amplitude of the wake. These competing processes can be modeled analytically for a satellite with the help of an empirical formula for the massloss history. The analytical model agrees with more traditional numerical simulations of tidal massloss and dynamical friction. Rapid massloss due to increasing tides at smaller and smaller radius makes it less likely for streams or remnants of infalling satellites to intrude the inner halo (as the Sgr stream and \\omegacen) than to stay in the outer halo (as the Magellanic stream), hence any intermediate-mass central black holes of the satellites are also likely \"hung up\" at large distances as well. It is difficult for the satellites' black holes to come close enough to merge into the supermassive black hole in the center of the host potential unless the satellites started with (i) pericenters much smaller than the typical distances to present-day observed satellites and with (ii) central density much higher than in the often seen finite density cores of observed satellites. ", "introduction": "Current theory of galaxy formation favors the idea that galaxies form hierarchically by merging smaller lumps or satellites. The gravity from a moving satellite pulls behind it a \"wake\" of particles of the host galaxy (ChandraSekhar 1943, Mulder 1983). This dynamical friction dissipates the orbital energy of the satellites so that they sink deep into the host galaxy potential well, where they are disintegrated and virialized via baryonic feedbacks and tidal stripping. These processes might have determined the density profile of virialized halo of the host galaxy (Syer \\& White 1998, Dekel et al. 2003). There are about 150 and 300 globular clusters, and a few dozen dwarf satellites of mass $10^{6-9}\\msun$ in the Milky Way and M31 respectively. It is tempting to associate these dwarfs satellites and globular clusters as the markers/remnants of past hierarchical merging events. Indeed there are several examples of possible streams of remnants in the Milky Way (Lynden-Bell \\& Lynden-Bell 1995) including the recently found Galactic ring or Carnis Major dwarf galaxy, traced by a grouping of globular clusters (Martin et al. 2004). A giant stream is found in the Andromeda galaxy (McConnachie et al. 2003, Ferguson et al, 2002). Among these the Sagittarius dwarf galaxy stream (Ibata et al. 1997) between radius 10-50 kpc from the Galactic center is perhaps the best example. It brings in at least 5 globular clusters to the inner halo, including M54, the 2nd most massive cluster of the Milky Way. A still mysterious object is \\omegacen, with about $10^6\\lsun$ at 5 kpc from the Galactic center which has the morphology of a globular cluster, but has multiple epoches of star formation and chemical enrichment (see Gnedin et al. 2002 and the \\omegacen symposium). Another example is the G1 cluster, a very massive globular-like object with about $10^6\\lsun$ at about 40 kpc from the M31 center (Meylan et al. 2001). A system (NGC1023-13) almost identical to G1 is also found in the S0 galaxy NGC1023, at a projected distance of about 40 kpc from the host galaxy (Larsen 2001). Freeman (1993) suggested that such systems are the remnants of nucleated dwarf satellites, with their outer tenuous dark matter and stars being removed by galaxy tides. M31 also has an unusual collection of clusters as luminous as \\omegacen within 5 kpc in projection from its double-peaked center. Given the above evidences or signs for infalling objects in our galaxy and M31, it is interesting to ask whether some of the inner globulars could have also been the result of mergers. Beyond the Local Group, minor mergers are sometimes speculated as the mechanism to deliver massive black holes and gas material into the nucleus of an AGN to account for the directions of the jets and nuclear dusty disks (Kendall, Magorrian, \\& Pringle 2003). A very interesting related issue is whether giant black holes acquire part of the mass by merging the smaller black holes in the nuclei of infalling satellites. For this paper we revisit the basic theoretical questions: what is the condition for a dwarf galaxy to decay into the inner halo? What are the possible outcomes of tidal stripping of a dwarf satellite? How often do we get a system like \\omegacen or a naked black hole near the host center? The answers to these questions will help us to test the validity of the theory of hierarchical merger formation of galaxies. The key mechanism for satellites to enter the inner galaxy is dynamical friction, where the gravity of the satellite creates a wake of overdensities in the particle distribution of the host galaxy, which in turn drags the motion of the satellite with a force proportional to $m(t)^2$, where $m(t)$ is the mass of the satellite. Another process is tidal disruption, where the object sheds mass with each pericenter passage, and the remnants are littered along the orbit of the satellite. The above two processes compete with and regulate each other: orbital decay increases the tidal field, which reduces the mass of the satellite, hence slows down the orbital decay. Some examples of these effects have been shown in Zhao (2002) in the case of \\omegacen. The analytical formula of ChandraSekhar is widely used for gaining insights on dynamical friction because of the time-consuming nature of the more rigorous N-body numerical simulation approach. It is a customary practice in previous works to model the orbital decay of a satellite as a point mass of a fixed mass. However, {\\it the fixed-mass approximation is invalid} and could seriously overestimate dynamical friction because of neglecting massloss ${dm(t) \\over dt}$. It is essential in calculations of satellite orbits to model the dynamical friction and massloss together since they regulate each other. In the past the massloss and the orbital decay are often modeled in the {\\it ab initio} fashion, resulting coupled non-linear equations without simple analytical solutions. In such models, the satellite mass is often modeled as a function of the satellite's tidal radius, hence various factors come in, including the orbital position of the satellite, the density profile of the satellite (e.g., Jiang \\& Binney 2000, Zhao 2002, Mouri \\& Taniguchi 2003, Kendall et al. 2003). However, these complications are not always necessary since the massloss history is rather similar in simulations with very different initial conditions (the mass is generally a stair-case like a function of the time), so could be parametrized in an empirical fashion, by-passing the uncertain assumptions of the satellite initial profile. This could be useful for exploring a large parameter space of the satellite initial condition. Another invalid approximation but common practice is to use a static potential for the host galaxy. This again is unphysical since galaxy halos do grow in hierarchical formation scenario partly because of galaxy merging, and partly because of the adiabatic contraction of the baryonic disk and bulge; galaxy rotation curves $V_{\\rm cir}(r,t)$ can change by significantly before and after the formation of baryonic disks and bulges in mass models for the Milky Way and M31 (Klypin, Zhao \\& Somerville 2002) and in generic CDM simulations with baryons (e.g. Wright 2003). The growing gravitational force tends to restrain the radial excursions of the satellite while preserving the angular momentum. The dynamical friction or drag force is also proportional to the growing density $\\rho(r,t)$ of ambient stars and dark particles in the host galaxy. In fact, it is conceptually simple to incorporate massloss and growing potential while keeping the problem analytically tractable: the deceleration $-{dv \\over dt}$ is simply proportional to the satellite mass $m(t)$ and the ambient density $\\rho(r,t)$ of the host galaxy at the time $t$. Without massloss ChandraSekhar's formula would predict very efficient braking of the orbits, enough to make a high-mass satellite of $\\sim 10^{10}\\msun$ (the mass of the LMC or M33 sized object) to decay from a circular orbit at $\\sim 100$kpc to the very center of a high brightness galaxy in a Hubble time, delivering remnants into the inner galaxy. Here we study the effect of massloss and growing potential on the result of orbital decay, and the distribution of remnants. We present fully analytical results for calculating the decay rate for satellites on eccentric orbits in a scale-free growing isothermal potential. The structure of the paper is following: S2 gives the analytical formulation of the problem, S3 presents results of application to globulars and dwarf satellites, S4 studies the relation between massloss and the satellite density profile. S5 discusses the progenitors of Sgr and \\omegacen, and S6 summarizes. ", "conclusions": "\\subsection{Effects of disk, bulge and orbital inclination} One limitation of the current analysis is that we assume an isothermal dark matter plus stars model throughout the galaxy, hence the dynamical friction effect of the disk and bulge are not modeled accurately. However, the disk and bulge are not important for our conclusion because we predict mainly the orbital decay in the outer halo where $j>j_{\\rm disk}=3000{\\rm kpc~km/s}$. Inside 15 kpc, our estimation of dynamical friction by an SIS model is inaccurate only for satellites on low inclination orbits. If satellites come in random inclinations, it is more common to find high inclination orbits, for which our models should be fairly accurate even inside 15 kpc. \\subsection{Effects of escaping stars} In the part of our formulation where we derive the satellite mass profile, we assume a simplifying static picture that the satellite's mass is peeled off in successive layers at the shrinking tidal radius. The picture in N-body simulations is more complicated, since satellite particles at all radii, e.g., the center of the satellite, could in principle be escaping at any time. So eq. (26) sets only a lower limit on the initial density of the satellite. A more rigorous model should making this correction, e.g., by introducing an empirical factor to correct this as in Jiang \\& Binney (2000). \\subsection{Possible orbits of progenitors of the Sgr stream} The Sgr dwarf and the Canis Major dwarf are the closest known dwarf galaxies, about 15 kpc from the center of the Milky Way and at the edge of the Milky Way disk. The Canis Major is on a (direct or retrograde) orbit slightly inclined from the plane of the Milky Way, and the Sgr is on a nearly polar orbit. Both orbits have a fairly low angular momentum with $S \\sim 20\\kpc$; the data on Sgr are more complete, and show that it oscillates between 10 kpc pericenter and 50 kpc apocenter. Both contain several globular clusters. It is possible that the two dwarfs are the stripped-down version of a more massive object, which has dynamically decayed from the outer halo. Interestingly a possible extension of the Sgr has been reported recently in the SDSS data near the position of the outer halo globular cluster NGC2419 (Newberg et al. 2003). There is a stream-like enhancement of halo A-colored stars at the SDSS magnitude of $g_0=20.3$ in the plane of the Sgr's orbit, corresponding to a distance of 90 kpc. If this is true, it would imply that the Sgr has changed its orbits in the past Hubble time. There are two possible ways that this could happen. One is that the Sgr's orbit has been deflected by a massive satellite, such as the LMC or SMC. Indeed the orbits of the Sgr and the Magellanic Clouds do overlap at the Galactic poles, and simple timing arguments show that these systems encounter or fly by each other about 2.5 Gyrs ago at about 50 kpc on the North Galactic Pole if the rotation curve of the Milky Way is nearly flat (Zhao 1998). The problem of this solution is that it is rare for the Sgr to receive a strong enough deflection to bring down its orbit. Another solution is that the Sgr has been a more massive system, which orbital decayed from the outer halo (Jiang \\& Binney 2000). {\\it Our Model B illustrates such an example of the progenitor of the Sgr} which had an initial mass comparable to the LMC ($10^{10}\\msun$) and was on an eccentric orbit with radius between $20-140\\kpc$ (cf Fig.6). This model is similar to the Model K of Jiang \\& Binney. After a Hubble time the orbit decays to a small orbit very much like that of the Sgr with peri-to-apo ratio of $10\\kpc:50\\kpc$. Large amount of the material of the progenitor is shed in the radius between $10-140\\kpc$, the stream near NGC2419 at 90 kpc could be part of this debris near one of the apocenters of the orbit. Unfortunately the present model ends with a mass of $5\\times 10^9\\msun$, too large for the present-day Sgr. Some fine tuning of initial conditions and detailed N-body simulations are clearly needed to test this idea. \\subsection{Possible orbits of the progenitor of \\omegacen} We have mainly concentrated on the problem of getting rid of a satellite's angular momentum if it starts with a high angular momentum or orbital size $S_0\\gg 15$kpc. What would be the remnant distribution if a satellite is born with an initial orbital size $S_i<15$kpc? The stars in such a system are assembled in the inner halo from the start, e.g., by colliding an infalling gas cloud with the protogalactic gas clouds in the inner halo, (Fellhauer \\& Kroupa 2002). Or the stars form from extragalactic gas and descend on a very radial orbit, penetrating the inner 15 kpc of the host halo from its very first pericentric passage. An intriguing example is \\omegacen. Unfortunately our analytical model is not suited for this system because it is presently on a low-inclination eccentric retrograde orbits between 1 and 6 kpc from the Galactic center (Dinescu et al. 1999), so the contribution of dynamical friction by the disk is important. Also hydrodynamical friction with the disk gas can play a role for an early-on partially gaseous satellite. Nevertheless, if one applies simplisticly the tidal massloss and ChandraSekhar's dynamical friction in a spherical halo, one finds that while it seems easy to peel off a satellite galaxy to make a central star cluster, most simulations produce remnants on much larger orbits than \\omegacen (Zhao 2002). It seems some fine tuning is required to select progenitors on very low angular momentum and/or low energy orbits: the initial angular momentum needs to be low enough for the progenitor to penetrate into the inner halo or the present location of \\omegacen on its very first pericentric passage. This means that the initial orbital size $S_i^\\omega$ of \\omegacen is in between the present value of \\omegacen $S_0 \\sim 1.25$kpc and the boundary of the inner halo $R_{\\rm disk}=15$kpc, or mathematically \\beq 1.25\\kpc < S_i^\\omega < 15\\kpc. \\eeq Most recently there have been several very encouraging attempts to model the dynamical and star formation history of \\omegacen by nearly self-consistent N-body simulations (Mizutani et al. 2003, Tsuchiya et al. 2003, Bekki \\& Freeman 2003). All are able to produce both a reasonable mass and orbit of the \\omegacen after some trial and error with the initial parameters of the progenitor; many initial conditions lead to remnants, unlike \\omegacen, beyond 10kpc of the Milky Way center. The favored initial orbit has a small orbital size $S_i$. according to Tsuchiya et al. $j_i=60 {\\rm kpc} \\times 20 \\kms =1200{\\rm kpc}\\kms$ (or $S_i=6$kpc) and according to Bekki \\& Freeman $j_i=25 {\\rm kpc} \\times 60 \\kms =1500{\\rm kpc}\\kms$ (or $S_i=7.5$kpc). The small orbital size seems consistent with our expectation (cf. Fig.3). Tsuchiya et al. launch satellites with various initial mass $(0.4-1.6)\\times 10^{10}\\msun$ and with either a King profile or a Hernquist profile from 60 kpc from the Milky Way center. They choose well-aimed nearly radial orbits, with an initial perigalactic radius about 1 kpc, much more radial than the present eccentric orbit. Massloss in their King model are similar to our exponential massloss models ($n=\\infty$): rapid in the beginning, and $\\log(m)$ is roughly linear with time up to a mass of $10^8\\msun$ when the satellite has too little mass to proceed with the orbital decay. Massloss in the Hernquist model is closer to a $n=0.3$ model, linear in the beginning and rapid just before complete disruption (cf Fig.1a). Our comparison with Tsuchiya et al.'s numerical model would be fair apart from one theoretical concern. The progenitor in their best simulation is a two-component \"nucleated\" model with a rigid nucleus modeled by a extended-particle of $10^7\\msun$ with a half mass radius of 35pc on top of a live satellite of $0.8\\times 10^{10}\\msun$ with a Hernquist profile of half-mass radius $1.4$kpc; the dynamical friction of the Hernquist halo helps to deliver the nucleus eventually to an orbit similar to that of \\omegacen. However, a closer examination reveals a subtle inconsistency in making the nucleus rigid: the tidal force from the Hernquist halo beats the self-gravity of this fluffy nucleus at its half-mass radius by a factor of a few, so it could not have stayed and being moved as one piece. Nevertheless, one could have used a more contact, hence more plausible model of the rigid nucleus, say, with a total mass of $3\\times 10^6\\msun$ and a smaller half-mass radius of 7pc, which are closer to the observed mass and half-mass radius of \\omegacen. With this in mind, the formal inconsistency in Tsuchiya et al.'s best simulation seems to be harmless, and their model shows that \\omegacen could in principle be the remnant of a massive satellite on an orbit of initial apo-to-peri ratio of (about) 60kpc/2kpc." }, "0403/astro-ph0403316_arXiv.txt": { "abstract": "We consider the stability of a configuration consisting of a vertical magnetic field in a planar flow on elliptical streamlines in ideal hydromagnetics. In the absence of a magnetic field the elliptical flow is universally unstable (the ``elliptical instability''). We find this universal instability persists in the presence of magnetic fields of arbitrary strength, although the growthrate decreases somewhat. We also find further instabilities due to the presence of the magnetic field. One of these, a destabilization of Alfven waves, requires the magnetic parameter to exceed a certain critical value. A second, involving a mixing of hydrodynamic and magnetic modes, occurs for all magnetic-field strengths. These instabilities may be important in tidally distorted or otherwise elliptical disks. A disk of finite thickness is stable if the magnetic fieldstrength exceeds a critical value, similar to the fieldstrength which suppresses the magnetorotational instability. ", "introduction": "The problem of momentum transport in accretion disks is widely believed to require hydrodynamic or hydromagnetic turbulence for its resolution. The origin of this turbulence may be sought in the instability of laminar solutions of the equations of hydromagnetics, solutions that are compatible with the geometry of accretion disks. The recent history of these efforts has taken the form of first recognizing such an instability mechanism, and then trying to incorporate that mechanicsm into realistic disk models. The magnetorotational instability (MRI) mechanism, originally discovered by \\citet{vel} and \\citet{cha} and first applied to accretion disks in \\citet{bah}, is of this kind (see \\citet{mri} for a review). It appears in rotating, magnetized systems in which the specific angular momentum increases outward and in which the magnetic field is weak enough that rotational effects are not overwhelmed by magnetic tension. A second mechanism, that of the elliptical instability considered by \\citet{jg} and others \\citep{lpk,ryug, ryugv}, is also consistent with the accretion-disk setting. This instability mechanism has been reviewed by \\citet{rk02}. In the setting considered by Goodman et. al., it appears to require a secondary in order to enforce departure from rotational symmetry of the streamlines via a tidal potential. This is certainly appropriate for binary systems but it is likely that, even in the absence of a secondary, the laminar motion in the plane of the disk would not be accurately circular, so the elliptical-instability mechanism would appear to be a candidate of considerable generality. It does not require a magnetic field. One of the conclusions of the present paper is that it further persists in the presence of a magnetic field. In the idealized setting of the present problem, the latter may be of arbitrarily large strength. However, we also argue that in the setting of a disk geometry, there may indeed be a limit on the field strength. In this paper we therefore investigate the interaction of a vertical magnetic field with flow on elliptical streamlines, on the ground that both magnetic fields and noncircular streamlines are likely ingredients in accretion-disk settings. There are similarities with and differences from previous work on effect of magnetic fields on the elliptical instability \\citep{rk94}, which are discussed in \\S \\ref{discussion}. ", "conclusions": "\\label{discussion} We have explored the effect of a uniform, vertical magnetic field on the stability of planar, incompressible flow with elliptical streamlines in an unbounded medium, in the approximation of ideal magnetohydrodynamics. In the absence of magnetic fields, flows with elliptical streamlines having ellipticity parameter $\\epsilon$ [see equation (\\ref{equ:definitions})] are known to be unstable to perturbations with wavevectors that are transverse to the plane of the flow (the ``elliptical instability\"). Our first conclusion is that this elliptical instability persists in the presence of a vertical magnetic field: the latter decreases the maximum growthrate but fails to suppress the instability, no matter how large the magnetic-field parameter becomes. It can be compared with the conclusion of \\citet{rk94} that a toroidal magnetic field has a stabilizing influence. Kerswell's analysis holds for small $\\epsilon$ only and shows that the growthrate decreases with magnetic field in that limit. Our result, which holds for a vertical magnetic field, shows a similar trend for small values of the magnetic-field parameter, but this trend never results in complete stabilization of the elliptical instability with increasing magnetic field. A second conclusion is that there are further instabilities associated with the presence of the magnetic field. One of these, for which the eigenvector is a mixture of hydrodynamic and magnetic modes, occurs for all values of the magnetic field parameter. Another, for which the eigenvector is a combination of magnetic modes only, sets in for values of the magnetic-field parameter exceeding a certain threshold value ($\\eta > \\sqrt{3}$). In all three of these instabilities, for large magnetic fields, the wave vector makes only a small angle with the plane of the unperturbed flow, reflecting the familiar tendency for dynamics to become nearly 2-dimensional in a strong, well ordered magnetic field. This is reflected in Figure \\ref{mixedmode}, which shows that as $\\eta$ increases, the unstable wedges are pushed to smaller $\\mu$. Although the unstable fraction of the $(E,\\mu)$ plane decreases with increasing $\\eta$ (except for a very slight maximum at $\\eta\\sim$ 2.18, reflecting the onset of instability between magnetic modes), the separation between the unstable wedges also decreases. While the nonlinear evolution of the unstable system is beyond the scope of this work, the destabilization of a nearly continuous swath of parameter space may have consequences for the interactions between unstable modes. In all three cases, the maximum instability increment tends to $\\epsilon \\pi/2$, i.e., the maximum growthrate of the unstable modes tends, in dimensional units, to $\\epsilon\\Omega /4$. All three of these instabilities may be relevant in accretion-disk settings. In systems of finite thickness $H$, however, the instability is suppressed if the Alfv\\'en speed $v_A$ exceeds a critical value of order $\\Omega H$. Magnetorotational instabilities are quenched at approximately the same fieldstrength \\citep{mri}. The growth rate of magnetoelliptical instabilities is smaller than that of magnetorotational instabilities by a factor of order $\\epsilon$, and thus they are not necessarily the primary instability in magnetized disks. They may well play a secondary role by breaking up eddies or vortices generated by other mechanisms. Magnetoelliptic instabilities may also occur in the inner parts of barred galaxies, in which the gas flow is slightly elliptical and the magnetic field, at least in the Milky Way, has a vertical component \\citep{mor}. In such settings, the instabilities could be a source of turbulence, possibly affecting the mass supply to a central compact object." }, "0403/astro-ph0403120_arXiv.txt": { "abstract": "We review the results obtained with the Galactic center campaigns of the BeppoSAX Wide Field X-ray Cameras (WFCs). This pertains to the study of luminous low-mass X-ray binaries (LMXBs). When pointed at the Galactic center, the WFC field of view contains more than half of the Galactic LMXB population. The results exemplify the excellent WFC capability to detect brief X-ray transients. Firstly, the WFCs expanded the known population of Galactic thermonuclear X-ray bursters by 50\\%. At least half of all LMXBs are now established to burst and, thus, to contain a neutron star as compact accretor rather than a black hole candidate. We provide a complete list of all 76 currently known bursters, including the new case 1RXS J170854.4-321857. Secondly, the WFCs have uncovered a population of weak transients with peak luminosities up to $\\sim10^{37}$~\\lum\\ and durations from days to weeks. One is the first accretion-powered millisecond pulsar SAX~J1808.4-3658. Thirdly, the WFCs contributed considerably towards establishing that nearly all (12 out of 13) luminous low-mass X-ray binaries in Galactic globular clusters contain neutron stars rather than black holes. Thus, the neutron star to black hole ratio in clusters differs from that in the Galactic disk at a marginal confidence level of 97\\%. ", "introduction": "Bleeker \\cite{bl03} reviewed in general terms the prospects and results of the BeppoSAX Wide Field Camera instrument package (``WFCs''). The unique capability of matching a wide field of view of $40^{\\rm o}\\times40^{\\rm o}$ per each of two identical cameras with a good angular resolution of 5\\arcmin\\ \\cite{jag97} not only led to a revolution of gamma-ray burst research (e.g., \\cite{pir03,hei03}) but also to a serious advance of our knowledge on X-ray bursts and other transient emission processes in low-mass X-ray binaries (LMXBs). Much of the relevant data was acquired during the only dedicated BeppoSAX observation program involving the WFCs as prime instrument. This encompasses observations pointed at the Galactic center during semi-yearly visibility windows that lasted from mid February to mid April and from mid August to mid October. Table~\\ref{tab1} summarizes the twelve campaigns during the six-year BeppoSAX lifetime. The combined exposure time represents 8\\% of the total BeppoSAX exposure budget. The success of this program may be anticipated by the mere fact that the field of view of one WFC emcompasses more than half the Galactic LMXB population according to pre-WFC catalogs \\cite{par95}. Another important ingredient for the success is that nearly all observations were analyzed in a near to real-time fashion. This was possible thanks to the 24 hr per day, 7 days a week, manning of the BeppoSAX Science Operations center (which was also crucial to the success of the GRB program of BeppoSAX \\cite{pir03}) and the dedicated support by the 'duty scientists'. Tied to this program a dedicated target-of-opportunity program was in place for the BeppoSAX Narrow Field Instruments (NFIs) to follow up new transients or bursters. This program was triggered 14 times, with exposure times between 20 and 40 ksec. Whenever new transients or bursters were discovered, these were announced in IAU circulars. This happened on 20 occasions and triggered independent TOO programs on BeppoSAX, the Rossi X-ray Timing Explorer (RXTE), XMM-Newton, and ground-based radio and optical telescopes, illustrating a community service of the WFC program. \\begin{table}[!t] \\caption{WFC observation campaigns on the Galactic center. The effective exposure times are for the position of the Galactic center. Adapted from \\cite{ha03,we03}.\\label{tab1}} \\begin{tabular}{ccrr} \\hline \\multicolumn{2}{c}{Campaign} & \\multicolumn{1}{r}{\\# obs.} & \\multicolumn{1}{r}{$t_{\\mathrm{exp}}\\,(\\rm{ks})$} \\\\ \\hline 1996 & Aug.15--Oct.29 & 67 & 1017 \\\\ 1997 & Mar.02--Apr.26 & 21 & 654 \\\\ 1997 & Sep.06--Oct.12 & 13 & 302 \\\\ 1998 & Feb.11--Apr.11 & 17 & 551 \\\\ 1998 & Aug.22--Oct.23 & 10 & 410 \\\\ 1999 & Feb.14--Apr.11 & 14 & 470 \\\\ 1999 & Aug.24--Oct.17 & 24 & 801 \\\\ 2000 & Feb.18--Apr.07 & 21 & 633 \\\\ 2000 & Aug.22--Oct.16 & 29 & 767 \\\\ 2001 & Feb.14--Apr.23 & 5 & 215 \\\\ 2001 & Sep.04--Sep.30 & 7 & 284 \\\\ 2002 & Mar.05--Apr.15 & 5 & 91 \\\\ \\hline \\end{tabular} \\end{table} In the present paper, we provide a general overview of the accomplishments obtained through the WFC Galactic center program. We categorize the achievements in three areas: thermonuclear X-ray bursts, transients, and luminous globular cluster sources. ", "conclusions": "The results of the Galactic center campaigns of the BeppoSAX WFCs are plentiful. This is mainly due to its large field of view which enabled to monitor a large fraction of the Galactic LMXB population and resulted in a large exposure of roughly 7 million seconds over six years on tens of sources. Such an exposure has not been accomplished by any other device yet. Thus, it has been possible to detect rare phenomena such as superbursts, burst-only sources and swift transients (e.g., V4641 Sgr). The analysis of WFC data will continue and new results are expected to come out. This particularly concerns time scales of a few hours and the longest time scales, and weak signals (e.g., X-ray bursts fainter than 0.5~Crab)." }, "0403/astro-ph0403250_arXiv.txt": { "abstract": "The Chandra Multiwavelength Project (ChaMP) has discovered a jet-like structure associated with a newly recognized QSO at redshift z=1.866. The system was 9.4\\arcmin\\ off-axis during an observation of 3C 207. Although significantly distorted by the mirror PSF, we use both a raytrace and a nearby bright point source to show that the X-ray image must arise from some combination of point and extended sources, or else from a minimum of three distinct point sources. We favor the former situation, as three \\emph{unrelated} sources would have a small probability of occurring by chance in such a close alignment. We show that interpretation as a jet emitting X-rays via inverse Compton (IC) scattering on the cosmic microwave background (CMB) is plausible. This would be a surprising and unique discovery of a radio-quiet QSO with an X-ray jet, since we have obtained upper limits of 100 $\\mu$Jy on the QSO emission at 8.46 GHz, and limits of 200 $\\mu$Jy for emission from the putative jet. ", "introduction": "The objectives of the \\axaf\\ Multiwavelength Project (ChaMP) include identification and categorization of a complete, well-defined sample of serendipitous sources \\citep{Kim03, Green03}. The results will be of use, e.g., to study luminosity functions and their evolution, to quantify the newly resolved source(s) of the hard diffuse X-ray background, and to study cosmic structure and clustering of AGN and galaxies. The wide angle nature of this survey also makes it ideal to discover rare and unusual objects suitable for detailed study; e.g., lensed QSOs and X-ray jets. Schwartz (2002a,b) has pointed out that if the jets observed in X-rays on scales of tens to hundreds of kpc are emitting via IC scattering of the CMB as suggested by \\citet{Tavecchio00} and \\citet{Celotti01}, then they will maintain the same apparent surface brightness independent of redshift, and therefore can be detected to arbitrarily large redshifts, up to the epoch at which they form. The \\emph{Chandra} observations of such large scale jets in QSOs and powerful FR II radio sources are typically interpreted as IC/CMB emission, \\citep{Schwartz00, Harris02, Marshall01,Sambruna01,Siemiginowska02}. All such interpretations require the assumption that the jet is either relativistically beamed with Doppler factors of order $\\delta \\sim$ 3 to 15, or that the energy density in relativistic electrons grossly exceeds the magnetic field energy density by at least two orders of magnitude. Detection of the X-ray ``beacons'' predicted by Schwartz (2002a,b) would provide additional evidence that the above assumptions are well founded. We report the discovery of a candidate for such a system: CXOMP J084128.3+131107, (hereafter called J0841). The X-ray image shows an elongated structure. Despite the broad point response function (PSF) of the \\axaf\\ telescope at this 9.4\\arcmin\\ off-axis angle, we show that at least three point sources would be required to simulate the observed extent. We favor an interpretation of emission from the jet of an optically identified QSO which is close to the peak X-ray intensity. We also mention alternate interpretations. Due to the small probability for three \\emph{unrelated} sources to occur by chance in this configuration, such interpretations may be even more unusual. ", "conclusions": "\\citet{Schwartz02} has noted that X-ray emission by IC/CMB should result in X-ray jets being cosmic beacons -- maintaining the same surface brightness at any larger redshift. This is because the (1+z)$^{-4}$ cosmic diminution of surface brightness is exactly compensated by the (1+z)$^{4}$ increase in the energy density of the CMB with redshift. Such an effect does not depend on equipartition, or on relativistic beaming. The low magnetic field, $\\le$2 $\\mu$G, implied by the limits to radio emission is unusual. Fields in clusters of galaxies can approach 1 $\\mu$Gauss, while typical jet fields on kpc scales are of order 10 $\\mu$Gauss. So the upper limits to magnetic field strengths derived here are somewhat weak for a jet. However, there seems to be no fundamental physics prohibiting massive black holes to produce jets of such low internal energy density. Selection bias against finding radio quiet X-ray jets could explain why such low magnetic field jets have not previously been noticed. Alternately, this object may have a magnetic field much weaker than the equipartition value." }, "0403/astro-ph0403299_arXiv.txt": { "abstract": "Using evolutionary population synthesis (EPS) we present integrated spectral energy distributions (ISEDs) and absorption-line indices defined by the Lick Observatory image dissector scanner (referred to as Lick/IDS) system, for an extensive set of instantaneous burst single stellar populations (SSPs). The ages of the SSPs are in the range $1 \\leq \\tau /{\\rm Gyr} \\leq 19$ and the metallicities $-2.3 \\leq {\\rm [Fe/H]} \\leq +0.2$. Our models use the rapid single stellar evolution (SSE) algorithm of \\citet*{hur2000} for the stellar evolutionary tracks, the empirical and semi-empirical calibrated BaSeL-2.0 model of \\citet*{lej97b,lej98} for the library of stellar spectra and empirical fitting functions of \\citet{wor94b} for the Lick/IDS spectral absorption feature indices. Applying our synthetic Lick/IDS absorption-line indices to the merit function we obtain the age and the metallicity of the central region of M32, it can be well interpreted with an instantaneous SSP with an age of $\\sim$ 6.5\\,Gyr and a metallicity similar to solar. Applying the derived age and the metallicity from the merit function to a number of index-index diagrams, we find that the plots of $H_\\beta - Fe5015$ and $H_\\beta - Fe5782$ are the best index-index diagrams from which we can directly obtain reasonable age and metallicity. ", "introduction": "Age effects often mask metallicity effects in the studies of stellar populations \\citep{oco76,oco86,oco94,wor92}, and separating them is a cumbersome affair \\citep*{ren86,buz92,buz93}. The age-metallicity degeneration originates from the fact that increasing either the metallicity or the age makes the integrated spectral energy distribution (ISED) of a single stellar population (SSP) redder \\citep*{bre96}. Previous studies showed that it is difficult to break this degeneration only by broad-band colours \\citep{ari96,wor94a}. In order to solve this question, some studies used spectral information, instead of colours only, in the evolutionary population synthesis (EPS) models \\citep*{bre94,bru93,bru96,bru2003,buz89,kod97,tan96,vaz99a,wor94a}. With the development of these models including spectral information, the spectral resolution has been improved from about 20\\,\\AA \\ to 2\\,\\AA \\ \\citep{vaz99a}. Furthermore, these spectral information has been translated to the line strengths either by empirical fitting polynomials \\citep{gor93,wor94b,wor97} or approaches other than the Lick fitting functions \\citep{pel89,buz95}. The inclusion of absorption feature indices in EPS models could add the power of diagnostics to the study of stellar populations \\citep{wor94a}, therefore some studies have included spectral absorption feature indices into EPS models attempting to understand the stellar populations. Some earlier EPS studies contained spectral indices include no metallicity dependence \\citep{bru83,tin72a,tin72b,tin76} or are otherwise limited in scope \\citep*{aar78,fro80,mou78,tri92}. Recently, the spectral absorption indices have been combined systematically in EPS studies, in those models some model builders use the absorption feature indices at intermediate resolution (9\\,\\AA) of Lick system \\citep*{bru96,bru2003,jon95,kur99,pel89,vaz96,vaz99a,wor94a}, which has been accepted by many investigators and widely used in their studies; other model builders use the indices at considerably higher resolution (FWHM $\\sim$ 2\\,\\AA) in Rose system \\citep{buz92,buz93,gor93,ros94,vaz99a,vaz2001}. SSPs are assemblies of chemically homogeneous and coeval single stars, the star formation history of any stellar system can be described by a superposition of SSP models with different ages and metallicities. Studying SSPs can help us to understand the evolution of clusters and galaxies, the distribution of metallicities, and to quantify the star formation history of the galaxies. In this paper we also investigate SSPs with a systematic and self-consistent set of stellar evolution models by \\citet{pol98}, spectral library -- the BaSeL-2.0 model \\citep{lej97a,lej97b,lej98} and the spectral absorption feature indices of Lick system \\citep{gor93,wor94b}. The outline of the paper is as follows. In Section 2, we describe our SSPs models and algorithm. In Section 3, we give the results and discussion. In Section 4, we apply the observed spectral indices to the merit function to determine the age and the metallicity for SSP-like assembly. Finally we present the summary and conclusions in Section 5. ", "conclusions": "In this paper we use the EPS technique to present the ISEDs at intermediate resolution (10\\,\\AA \\ in the ultraviolet and 20\\,\\AA \\ in the visible) and the Lick/IDS spectral absorption indices for an extensive set of instantaneous burst SSPs of various ages and metallicities ($1 \\leq \\tau \\leq 19\\,$Gyr, $-2.3 \\leq {\\rm [Fe/H]} \\leq 0.2$). In our EPS models we adopted the rapid SSE algorithm, the empirical and semi-empirical calibrated BaSeL-2.0 spectral library and the fitting functions for the absorption-line indices. As checks on our models, we compare our ISEDs with those of Worthey and the high resolution spectrum of NGC 6838, also compare the Lick/IDS absorption feature indices with the values of other EPS models and those of Galactic and M31 globular clusters and the central region of M32. We conclude that: (a) The disagreement between our ISEDs with those of Worthey presents in two wavelength ranges, one in the far-ultraviolet region, the other in the visible and infra-red regions. The discrepancy of ISEDs in the far-ultraviolet region is mainly caused by the inclusion of hotter PN stars in our models, while the discrepancy of ISEDs in the visible and infra-red regions is caused by the choice of stellar evolutionary models and spectral library. The inclusion of cooler TPAGB/PPN stars can produce a significant deviation in ISEDs for young SSPs, but almost no effect for old SSPs in the visible and infra-red regions. (b) The integrated spectrum of NGC 6838 can be fitted by a SSP with solar metallicity and an age of $\\sim$ 12.6\\,Gyr. (c) When comparing our synthetic Lick/IDS absorption-line indices with the values in the literatures and observational data, we find that our models are generally in agreement with other models and match real globular clusters in most index-index diagrams. $TiO_{\\rm 1}$ is, however, systematically smaller. At last using the merit function $F$, we find the that spectral indices of the central region of M32 can be fitted with a single stellar population with an age of $\\sim$ 6.5\\,Gyr and a solar-like metallicity with $F=1.00$. Applying the age and the metallicity of M32 to some index-index diagrams, we see that $H_\\beta - Fe5015$ and $H_\\beta - Fe5782$ are better diagrams to estimate directly the age and metallicity from index-index plots." }, "0403/astro-ph0403585_arXiv.txt": { "abstract": "We analyse near-infrared {\\em HST}/NICMOS $F110W (J)$ and $F160W (H)$ band photometry of a sample of 27 $i'$-drop candidate $z\\simeq6$ galaxies in the central region of the HST/ACS Ultra Deep Field (HUDF). The infrared colours of the 20 objects not affected by near neighbours are consistent with a high redshift interpretation. This suggests that the low redshift contamination of this $i'$-drop sample is smaller than that observed at brighter magnitudes where values of 10-40\\% have been reported. The $J-H$ colours are consistent with a slope flat in $f_\\nu$ ($f_\\lambda\\propto\\lambda^{-2}$), as would be expected for an unreddened starburst. There is, however, evidence for a marginally bluer spectral slope ($f_\\lambda\\propto\\lambda^{-2.2}$) which is perhaps indicative of an extremely young starburst ($\\sim10$ Myr old) or a top heavy initial mass function and little dust. The low levels of contamination, median photometric redshift of $z\\sim6.0$ and blue spectral slope, inferred using the near-infrared data, supports the validity of the assumptions in our earlier work in estimating the star formation rates and, that the majority of the i-drop candidates galaxies lie at $z\\sim6$. ", "introduction": "\\label{sec:intro} In recent years the observational horizon has expanded rapidly and radically for those observing distant galaxies. Large format red-sensitive detectors on wide field imaging instruments, the new generation of 8m class telescopes and the refurbished Hubble Space Telescope ({\\em HST}), have pushed the limits to which we can routinely detect star-forming distant galaxies progressively from redshifts of one to beyond $z=6$. At the highest redshifts currently accessible, narrow band emission lines searches using the Lyman-$\\alpha$ line have moved on from redshifts of 4 \\citep{hm96,fu03}, to 5.7 \\citep{hu99} and now reach to $z\\sim6.5$ \\citep{ko03,hu02}. Photometric redshifts \\citep[e.g.][]{lan96,co97,fe99} are now routinely used to process large datasets and identify high redshift candidates. An application of this method, the continuum based Lyman-break photometric technique pioneered at $z\\sim3$ by \\citep{guh90,st95}, has been extended progressively to z$\\sim4$ and $z\\sim5$ \\citep[e.g.][]{st99,br04}, spectroscopically confirmed, and using $i'$-band drop selection further extended to $z\\sim6$ \\citep{st03,st04a,st04b,bo03a,bb03,bo04,di04,yan03}. The `classical' Lyman break technique used by Steidel et al.\\ (1996) uses three filters, one redward of rest frame Lyman-$\\alpha$ ($\\lambda_\\mathrm{rest}>1216$\\AA), a second in the spectral region between rest-frame Lyman-$\\alpha$ and the rest-frame Lyman limit (912\\AA) and a third at $\\lambda_\\mathrm{rest}<912$\\AA. At z$\\sim$3 the technique relies on the ubiquitous step in the spectra of the stellar component of galaxies at 912\\AA\\ due to photospheric absorption, supplemented by optically-thick Lyman limit system absorption caused by neutral hydrogen in the galaxy in question or in the intervening IGM. At higher redshifts the evolution in the Lyman-$\\alpha$ forest absorption, particularly in the spectral region 912-1216\\AA\\ means that the effective break migrates redward to the Lyman-$\\alpha$ region. Recently this search technique has been extended to $z\\sim6$ by various authors and $i'$-drop samples have been used to constrain the star formation history of the universe (as derived from rest-frame UV luminous, starbursting stellar populations, e.g. Stanway, Bunker \\& McMahon 2003; Giavalisco et al.\\ 2004; Bouwens et al.\\ 2004; Bunker et al.\\ 2004). Any such estimation, however, is dependent on the characteristic properties of the $i'$-drop population. Although Steidel and coworkers \\citep{st99,st01} have successfully shown that a Lyman-break technique cleanly selects galaxies at intermediate redshift ($z\\approx3-4$), the $i'$-drop varient of the technique lacks a second colour to constrain the redshift distribution and is subject to contamination from both low redshift elliptical galaxies and cool dwarf stars. Until recently, these factors have been essentially unconstrained by observational data. The {\\em HST}/ACS Ultra Deep Field (HUDF, Beckwith, Somerville and Stiavelli 2004), imaging an 11 arcmin$^2$ region of the sky to faint magnitudes in the $F435W(B)$, $F606W(v)$, $F775W(i')$ and $F850LP(z')$ bands, has now allowed a clean sample of $i'$-drop objects to be defined and constraints to be placed upon its luminosity function \\citep{b04}. The {\\em HST}/NICMOS treasury programme, complimentary to the HUDF, has imaged the central region of the ACS field to faint magnitudes at wavelengths of 1.1 and 1.6 microns, allowing the infrared properties of the $i'$-drop population (in particular their rest-frame ultra-violet spectral slope, faint-end contamination and luminosity at 1500\\AA) to be determined. In this paper we discuss the infrared properties of the $i'$-drop population defined in \\citet{b04}, and their implications for the nature of these objects. We adopt a $\\Lambda$-dominated, `concordance' cosmology with $\\Omega_{\\Lambda}=0.7$, $\\Omega_{M}=0.3$ and $H_{0}=70\\,h_{70} {\\rm km\\,s}^{-1}\\,{\\rm Mpc}^{-1}$. All magnitudes in this paper are quoted in the AB system \\citep{og83} and the \\citet{ma95} prescription, extended to $z=7$, is used where necessary to estimate absorption due to the intergalactic medium. Beyond $z=7$ absorption is assumed to decrease the transmitted flux by 98\\% \\citep[the decrement observed in the flux of $z>6$ quasars identified by the Sloan Digital Sky Survey, ][]{fa03} . ", "conclusions": "\\label{sec:conc} \\begin{enumerate} \\item The near-infrared properties of a faint sample of $i'$-drop objects (defined by Bunker et al.\\ 2004) has been investigated, using the deep $F110W$ and $F160W$ images of the HST/ACS Ultra Deep Field. \\item The infrared colours of all objects not suffering from contamination by near neighbours are consistent with a high redshift interpretation. This suggests that the contamination of this $i'$-drop sample is smaller than that observed at brighter magnitudes. \\item The spectral slopes inferred from near infrared colours are consistent with $\\beta=-2.0$ although there is evidence for a marginally steeper spectral slope of $\\beta=-2.2\\pm0.2$. \\item These steep spectral slopes suggest that the dust extinction of the sample is small and that the $i'$-drop population may comprise galaxies with young starbursts - possibly triggered by mergers (several close systems of galaxies are observed in this sample). \\item The low levels of contamination, and the steep spectral slope, inferred from near-infrared data supports the validity of our previous star formation density estimates based on $i'$-drop samples. \\end{enumerate} \\subsection*{Acknowledgements} 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 programs \\#9803 and \\#9978. We thank Paul Hewett, Richard Ellis and Malcolm Bremer for useful discussions. ERS acknowledges a Particle Physics and Astronomy Research Council (PPARC) studentship supporting this study. We also thank Steve Beckwith and the HUDF team for making these high quality images publically available." }, "0403/astro-ph0403066_arXiv.txt": { "abstract": "In this paper we present a multiresolution-based method for period determination that is able to deal with unevenly sampled data. This method allows us to detect superimposed periodic signals with lower signal-to-noise ratios than in classical methods. This multiresolution-based method is a generalization of the wavelet-based method for period detection that is unable to deal with unevenly sampled data, presented by the authors in \\cite{otazu}. This new method is a useful tool for the analysis of real data and, in particular, for the analysis of astronomical data. ", "introduction": "\\label{intro} The detection of periodic signals in astronomical data has been usually addressed by classical Fourier-based or epoch folding methods. These methods have different problems when dealing with non-sinusoidal periodic signals or with very low signal-to-noise ratios. When the analysed data set contains several periodic signals, the behavior of classical period determination methods highly depends on intrinsic signal characteristics (see, for example, \\citealt{lafler}; \\citealt{jurkevich}; \\citealt{stellingwerf}; \\citealt{lomb}; \\citealt{scargleA}, \\citeyear{scargleB}; \\citealt*{roberts}; \\citealt{press}). The reader is referred to the introduction of \\cite{otazu} (hereafter Paper~I) for a general discussion. To avoid this problem, we presented there, as a preliminary step, a wavelet-based approach that only works with evenly time sampled data. In astronomy, however, this is not a usual situation, since data are mostly acquired on irregular intervals of time. In such a case there are two possibilities: resample the data into a new evenly sampled data set, or use a method able to deal with the original unevenly sampled data set. In the first case we are forced to modify the original data, which necessarily implies a loss of information. Moreover, this is not always possible if the temporal gaps are larger than some of the periods present in the data. In order to avoid these problems, a technique capable to deal with unevenly sampled data is needed. The present paper is a natural extension of Paper~I that allows to work on unevenly time sampled data. We show how the methodology of multiresolution decomposition (similar to the wavelet decomposition philosophy) is very well suited to this problem, since it is completely oriented towards decomposing functions into several frequential characteristics. As in Paper~I, the main objective is to isolate every signal present in our data and to analyse them separately, avoiding their mutual influences. In Section~\\ref{multiresolution} we outline some concepts in multiresolution analysis and their similarities with wavelet theory that are relevant to the stated problem. In Section~\\ref{period} we propose an algorithm to detect each of the periodic signals present in a data set by combining multiresolution analysis decomposition with classical period determination methods. In Sections~\\ref{simulated} and \\ref{results} we present some examples of synthetic data we used to test the algorithm and the results we obtained. We summarise our conclusions in Section~\\ref{conclusions}. ", "conclusions": "\\label{conclusions} In this paper we have presented a multiresolution-based method for period determination able to deal with unevenly sampled data. This constitutes a significant improvement with respect to the wavelet-based method presented in Paper~I, which is unable to deal with unevenly sampled data. The overall performance of the present method is similar to the wavelet-based one, in the sense that it allows us to detect superimposed periodic signals with lower signal-to-noise ratios than in classical methods. We stress that one advantage of the present method over classical methods is the simultaneous detection of a period in more than one multiresolution plane, allowing to improve the confidence of a given detection. Moreover, since here we are not forced to lose or modify the information when averaging or interpolating the original data, we can reach higher noise-to-signal ratios than in the wavelet-based method described in Paper~I. We note that the multiresolution decomposition scheme that we have used can be interpreted as a particular case of scale-space filtering. In order to improve isolation of periodic features, more general approaches could be used to perform this decomposition. In this context, anisotropic diffusion schemes proposed by \\cite{perona} could be useful if properly tuned." }, "0403/hep-ph0403003_arXiv.txt": { "abstract": "In a model of TeV right-handed (RH) neutrino by Krauss, Nasri, and Trodden, the sub-eV scale neutrino masses are generated via a 3-loop diagram with the vanishing see-saw mass forbidden by a discrete symmetry, and the TeV mass RH neutrino is simultaneously a novel candidate for the cold dark matter. However, we show that with a single RH neutrino it is not possible to generate two mass-square differences as required by the oscillation data. We extend the model by introducing one more TeV RH neutrino and show that it is possible to satisfy the oscillation pattern within the modified model. After studying in detail the constraints coming from the dark matter, lepton flavor violation and the muon anomalous magnetic moment, and the neutrinoless double beta decay, we explore the parameter space and derive predictions of the model. Finally, we study the production and decay signatures of the TeV RH neutrinos at TeV $e^+ e^-/\\mu^+ \\mu^-$ colliders. ", "introduction": "One of the most natural way to generate a small neutrino mass is via the see-saw mechanism \\cite{see-saw}. There are very heavy right-handed neutrinos, which are gauge singlets of the standard model (SM), and so they could have a large majorana mass $M_R$. After electroweak symmetry breaking, a Dirac mass term $M_D$ between the right-handed and the left-handed neutrinos can be developed. Therefore, after diagonalizing the neutrino mass matrix, a small majorana mass $\\sim m_D^2/M_R$ for the left-handed neutrino is obtained. This is a very natural mechanism, provided that $M_R \\sim 10^{11-13}$ GeV. One drawback of this scheme is that these right-handed neutrinos are too heavy to be produced at any terrestrial experiments. Therefore, phenomenologically there are not many channels to test the mechanism. Although it could be possible to get some hints from the neutrino masses and mixing, it is rather difficult to reconstruct the parameters of the right-handed neutrinos using the low energy data \\cite{kim}. Another natural way to generate a small neutrino mass is via higher loop processes, e.g., Zee model \\cite{zee}, with some lepton number violating couplings. However, these lepton number violating couplings are also subject to experimental constraints, e.g., $\\mu \\to e \\gamma, \\tau \\to e\\gamma$. In the Zee model, there are also extra scalars whose masses are of electroweak scale, and so can be observed at colliders \\cite{gl}. On the other hand, recent cosmological observations have established the concordance cosmological model where the present energy density consists of about $73\\%$ of cosmological constant (dark energy), $23\\%$ (non-baryonic) cold dark matter, and just $4\\%$ of baryons. To clarify the identity of the dark matter remains a prime open problem in cosmology and particle physics. Although quite a number of promising candidates have been proposed and investigated in detail, other possibilities can never be neglected. Recently, Krauss, Narsi, and Trodden \\cite{krauss} considered an extension to the SM, similar to the Zee model, with two additional charged scalar singlets and a TeV right-handed neutrino. They showed that with an additional discrete symmetry the Dirac mass term between the left-handed and right-handed neutrinos are forbidden and thus avoiding the see-saw mass. Furthermore, the neutrino mass can only be generated at three loop level, and sub-eV neutrino masses can be obtained with the masses of the charged scalars and the right-handed neutrino of order of TeV. Phenomenologically, this model is interesting because the TeV right-handed neutrino can be produced at colliders and could be a dark matter candidate. In this work, we explore in details the phenomenology of the TeV right-handed (RH) neutrinos. We shall extend the analysis to three families of left-handed neutrinos and explore the region of the parameters that can accommodate the present oscillation data. In the course of our study, we found that the model in Ref. \\cite{krauss} with a single RH neutrino cannot explain the oscillation data, because it only gives one mass-square difference. We extend the model by adding another TeV RH neutrino, which is slightly heavier than the first one. We demonstrate that it is possible to accommodate the oscillation pattern. We also obtain the relic density of the RH neutrino, and discuss the possibility of detecting them if they form a substantial fraction of the dark matter. We also study the lepton number violating processes and the muon anomalous magnetic moment, and the production at leptonic colliders. In particular, the pair production of $N_1 N_2, N_2 N_2$ at $e^+ e^-/\\mu^+\\mu^-$ colliders gives rise to very interesting signatures. The $N_2$ so produced will decay into $N_1$ plus a pair of charged leptons inside the detector. Thus, the signature would be either one or two pairs of charged leptons plus a large missing energy. The organization is as follows. We describe the model in the next section. In Sec. III, we explore all the phenomenology associated with the TeV RH neutrino. In Sec. IV, we discuss the signatures in collider experiments. Section V is devoted to the conclusion. ", "conclusions": "In this paper, we have discussed a model that explains the small neutrino mass and dark matter in the universe at the same time. Such a model was proposed by Krauss {\\it et al.} as a modification of Zee model. However, our study revealed that their original model is unfortunately not capable of explaining the neutrino oscillation pattern. We have extended the model by introducing another right-handed neutrino. We succeed in showing that such an extension is possible to achieve the correct neutrino mixing pattern. A prediction of this model is the normal mass hierarchy. In addition, the undiscovered mixing angle $\\theta_{13}$ is relatively large, because of the requirement of a mild cancellation between the parameters for a small $\\theta_{13}$ and a sensible coupling of the charged scalar, $\\lambda_s$. The relic density of the lightest right-handed neutrino has also been revisited. Under the constraint by WMAP we found that the mass of the right-handed neutrino cannot be as large as TeV but only of order $1\\times 10^2$ GeV, after a careful treatment of the calculation. In addition, other constrains including the muon anomalous magnetic moment, radiative decay of muon, neutrinoless double beta decay have also been studied. With all the constraints we are still able to find a sensible region of parameter space. Finally, our improved model has an interesting signature at leptonic colliders via pair production of right-handed neutrinos, in particular $N_1 N_2$ and $N_2 N_2$. The $N_2$ so produced will decay into $N_1$ plus two charged leptons. Thus, the signature is either one or two pairs of charged leptons with a large missing energy. Hence, this model can be tested not only by neutrino experiments but also by collider experiments." }, "0403/astro-ph0403700_arXiv.txt": { "abstract": " ", "introduction": "At the heart of a successful theory of galaxy formation must be a detailed physical understanding of the dissipational processes which form spiral galaxies. The disk is the defining stellar component of disk galaxies, and understanding its formation is in our view the most important goal of galaxy formation theory. Although much of the information about the pre-disk state of the baryons has been lost in the dissipative process, some tracers are likely to remain. What do we mean by the reconstruction of early galactic history? We seek a detailed physical understanding of the sequence of events which led to the Milky Way. Ideally, we would want to associate components of the Galaxy to elements of the protocloud -- the baryon reservoir which fueled the stars in the Galaxy. For many halo stars, and some outer bulge stars, this may be possible with phase space information. But for much of the bulge and the disk, secular processes cause the populations to become relaxed (i.e. the integrals of motion are partially randomized). In order to have any chance of unravelling disk formation, we must explore chemical signatures in the stellar spectrum. The exponential thin disk, with a vertical scale height of about 300 pc, is the most conspicuous component in edge-on disk galaxies. The thin disk is believed to be the end product of the quiescent dissipation of most of the baryons and contains almost all of the baryonic angular momentum. Many disk galaxies show a second fainter disk component with a longer scale height (typically about 1 kpc); this is known as the thick disk \\citep{sd00,db02}. The Milky Way has a significant thick disk \\citep{gr83}: its surface brightness is about 25\\% of the thin disk's surface brightness, and its stars are significantly more metal poor than the stars of the thin disk. The galactic thick disk is currently believed to arise from heating of the early stellar disk by one or more accretion events although other possible origins have been discussed \\citep{gwk89,gwj95}. It is composed of only old stars, with ages greater than 10 Gyr, equivalent to forming at a redshift of $z\\geq1$ \\citep{rw00}. Furthermore, the thick disk is chemically distinct from the thin disk at low metallicity \\citep{kf98,fbl03}, in the sense that [$\\alpha$/Fe] is enhanced relative to the thin disk, although solar values are seen at higher metallicity. The thick disk may be one of the most significant components for studying signatures of galaxy formation because it presents a 'snap frozen' relic of the state of the (heated) early disk. In the discussion which follows, we make reference to the KAOS project (see http://www.noao.edu/kaos) which proposes a highly multiplexed wide-field multi-object spectrograph for the Gemini Observatory. Full specifications for the proposed instrument can be found at the above web site. ", "conclusions": "" }, "0403/astro-ph0403536_arXiv.txt": { "abstract": "We combine the latest observations of disk galaxy photometry and rotation curves at moderate redshift from the FORS Deep Field (FDF) with simple models of chemical enrichment. Our method describes the buildup of the stellar component through infall of gas and allows for gas and metal outflows. In this framework, we keep a minimum number of constraints and we search a large volume of parameter space, looking for the models which best reproduce the photometric observations in the observed redshift range ($0.52.2$ and $\\leq 2.2$, respectively. The lines correspond to the colours of simple stellar populations from the latest models of Bruzual \\& Charlot (2003), for three ages (from top to bottom): 8, 1, and 0.1~Gyr. For each age two metallicities are considered: $Z_\\odot$ (solid) and $Z_\\odot /3$ (dashed). Notice the bluer colours -- which are most sensitive to young stars -- suggest young stellar ages, whereas the red/NIR colours imply older ages. } \\label{fig:CCDs} \\end{figure} The work on $113$ disk galaxies in the range $0.14)$ AGN as revealed by X-ray observations. We also discuss prospects for future \\hbox{X-ray} surveys with \\chandra, \\xmm, and upcoming missions (\\S4). Throughout this paper, we adopt $H_0=70$~km~s$^{-1}$ Mpc$^{-1}$, $\\Omega_{\\rm M}=0.3$, and $\\Omega_{\\Lambda}=0.7$ (flat cosmology). \\begin{figure}[t!] \\centering \\includegraphics[height=10cm]{fig01.ps} \\caption{Adaptively smoothed image of the 2~Ms \\hbox{CDF-N}, constructed from data in the \\hbox{0.5--2~keV} (red), \\hbox{2--4~keV} (green), and \\hbox{4--8~keV} (blue) bands. Nearly 600 sources are detected in the $\\approx 448$~arcmin$^2$ field. The regions covered by the HDF-N and GOODS-N surveys are denoted. Adapted from D.M. Alexander, F.E. Bauer, W.N. Brandt, et~al., 2003, AJ, 126, 539.} \\label{fig:1} \\end{figure} ", "conclusions": "" }, "0403/astro-ph0403193_arXiv.txt": { "abstract": "Recently, \\citet{z10} claimed to have discovered a galaxy at a redshift $z=10$, and identified a feature in its spectrum with a hydrogen Ly$\\alpha$ emission line. If this identification is correct, we show that the intergalactic medium (IGM) around the galaxy must be significantly ionized; otherwise, the damping wing of Ly$\\alpha$ absorption by the neutral IGM would have greatly suppressed the emission line. We find either that the large-scale region surrounding this galaxy must have been largely reionized by $z=10$ (with a neutral fraction $\\la 0.4$) or that the stars within the galaxy must be massive ($\\ga 100\\, M_\\odot$), and hence capable of producing a sufficiently large \\ion{H}{2} region around it. We generalize these conclusions and derive the maximum \\Lya line flux for a given UV continuum flux of galaxies prior to the epoch of reionization. ", "introduction": "\\citet{z10} have reported the discovery of a galaxy at a redshift $z=10$, which is gravitationally magnified by a factor $\\sim 25-100$ by the foreground galaxy cluster Abel 1835. The redshift identification is supported by the spectroscopic detection of an emission line at $1.338\\mu$m, the photometric signature of an absorption trough at shorter wavelengths, and the lensing geometry. Here, we examine the far-reaching implications of associating this spectral feature with the Ly$\\alpha$ emission line from a galaxy at $z=10$. In particular, we show that if hydrogen in the intergalactic medium (IGM) around this galaxy were neutral, then the damping wing of its \\Lya absorption would have greatly suppressed the \\Lya emission line of the galaxy. Our results provide a strong incentive for obtaining higher quality data on this galaxy. More generally, we derive the maximum \\Lya line flux for a given UV continuum flux of galaxies prior to the epoch of reionization. Future detection of bright \\Lya lines from high-redshift galaxies can be used to improve the current lower limit on the redshift of reionization. Throughout, we adopt values for cosmological parameters derived from recent observations of the cosmic microwave background \\citep{WMAP, uros}. For the contributions to the energy density, we assume ratios relative to the critical density of $\\Omm=0.27$, $\\Oml=0.73$, and $\\Omega_b=0.046$, for matter, vacuum energy, and baryons, respectively. We also choose a Hubble constant $H_0=72\\mbox{ km s}^{-1}\\mbox{Mpc}^{-1}$, and a primordial scale-invariant ($n=1$) power-law power spectrum with $\\sigma_8=0.9$, where $\\sigma_8$ is the root-mean-square amplitude of mass fluctuations in spheres of radius $8\\ h^{-1}$ Mpc. ", "conclusions": "Figure 1 shows that if the galaxy discovered by \\citet{z10} indeed has a redshift $z=10$, then either its stars are very massive ($\\ga 100M_\\odot$) and hence capable of creating a large \\ion{H}{2} region around it, or the large-scale IGM around it has already been mostly reionized (with a neutral fraction $\\la 0.4$). The ratio of the galaxy's UV continuum flux in two bands is consistent with the first possibility, but better determination of the continuum spectrum or detection of other recombination lines (Bromm, Kudritzki, \\& Loeb 2001; Tumlinson, Giroux, \\& Shull 2001; Oh, Haiman, \\& Rees 2001) is required to make a convincing case. For example, a stellar population with a top-heavy IMF will produce substantial flux in He II Ly$\\alpha$ ($303.9 $\\AA) and He II $\\lambda 1640$. A search for features in the spectrum of a redshift $z=10$ galaxy at $\\lambda_{\\rm obs}= 3343$\\AA\\ and $1.8\\mu$m would constrain the first interpretation. The discovery of other galaxies may test the second possibility, since the damping wing of a neutral IGM limits the maximum \\Lya line flux of any galaxy prior to reionization based on its UV continuum flux. The imprint of the damping wing is insensitive to a variety of complicating factors. Clumping of gas and its infall onto the galaxy have a minor effect, since the damping wing averages the IGM over large physical scales ($\\sim 1$ Mpc). Moreover, the characteristic size of the \\ion{H}{2} region around a bright galaxy at $z\\sim 10$ is an order of magnitude larger than the scale of its infall region. The characteristic peculiar velocity of galaxies at $z\\sim 10$ is an order of magnitude smaller than the Hubble velocity at the edge of their \\ion{H}{2} regions, and can only significantly change the effect of resonant absorption on the line profile. We have ignored resonant absorption from \\ion{H}{1} within the \\ion{H}{2} region since it would not suppress even the blue wing of the \\Lya emission line if the galaxy had a peculiar velocity (in the direction of recession) larger than the line width (thus allowing the line to appear nearly symmetric); such a peculiar velocity may not be improbable within small samples because it makes the line more easily observable and adds a selection bias. In any case, the addition of resonant absorption would only strengthen our limits on the \\Lya line flux before reionization. Dust extinction in the host galaxy would affect ionizing photons (and thus the \\Lya line flux) more than it would the continuum at longer wavelengths. Based on the observed UV continuum, we would have predicted a weaker line with the inclusion of extinction, and so our argument would have been stronger. We have considered the lowest (most probable) value within the range of lensing magnification factors inferred by \\citet{z10}. A higher value would again strengthen our limits. For example, a magnification factor of 100 would imply that the line was 4 times weaker intrinsically, but the predicted line from the UV continuum would then be weakened by more than a factor of 4 since the smaller \\ion{H}{2} region would imply a damping wing that suppressed more of the line flux. Current observational constraints on reionization provide an inconsistent picture. On the one hand, the large-scale polarization anisotropies of the cosmic microwave background measured by WMAP imply a reionization redshift of 10-20 (Kogut et al.\\ 2003), while on the other hand, the extent of the \\ion{H}{2} regions around the highest redshift quasars indicates\\footnote{The evidence for a possible Gunn-Peterson (1965) trough in these quasar spectra (White et al. 2003) is still being debated (Songaila 2004).} a significantly neutral IGM at $z\\sim 6.4$ (Wyithe \\& Loeb 2004). Moreover, the IGM should have been cooler than observed at $z\\sim 3$-4 if hydrogen had been fully reionized at $z\\ga 9$ (Theuns et al.\\ 2002; Hui \\& Haiman 2003). Theoretically, it is possible that the ionization fraction evolved in a complex, non-monotonic fashion owing to an early episode of Pop III star formation (Wyithe \\& Loeb 2003; Cen 2003). Existing observations do not enable us to discriminate between the two interpretations of a positive detection of a \\Lya emission line from the $z=10$ galaxy of \\citet{z10}. Sokasian et al.\\ (2003a,b) have studied the process of reionization using cosmological simulations with radiative transfer; they have shown that for a Scalo-type IMF, reionization occurs over an extended redshift interval. For models in which the \\Lya optical depth at $z\\approx 6$ matches the value inferred from the SDSS quasars, a substantial fraction of the mass in the universe ($>30 \\%$) is already completely ionized by $z=10$. Since reionization proceeds ``inside-out'' affecting overdense regions first within the simulated volume (see also Gnedin 2000), we would expect the environment of the \\citet{z10} galaxy to be ionized. If the evolution had been more complex and included sources with population III stars prior to $z=12$, an even larger fraction of the mass would have been ionized by $z=10$ (Wyithe \\& Loeb 2003; Sokasian et al.\\ 2003b). This picture could change, however, if reionization occurred ``outside-in'' (e.g., Miralda-Escud\\'e, Haehnelt \\& Rees 2000). As shown by Barkana \\& Loeb (2004), the fluctuations in the real universe should be far larger than indicated in existing simulations because of their limited box size. In reality, small fluctuations on scales larger than the box are greatly amplified through biased galaxy formation at high redshifts. The various scenarios for reionization can be tested with more data on \\Lya emitting galaxies (supplementing recent work by Hu et al. 2002, 2004; Malhotra \\& Rhoads 2002; Kodaira et al.\\ 2003; Santos et al.\\ 2003; Kneib et al.\\ 2004; Stanway et al.\\ 2004; Barton et al.\\ 2004), SDSS quasars (Fan et al.\\ 2003), or gamma-ray burst afterglows (Loeb 2003; Barkana \\& Loeb 2004a). Over the next decade, it may also be possible to map directly the neutral hydrogen in the IGM through its $21\\,$cm line (see recent discussions by Zaldarriaga, Furlanetto, \\& Hernquist 2003; Morales \\& Hewitt 2003; Loeb \\& Zaldarriaga 2003; Gnedin \\& Shaver 2003; Sokasian, Furlanetto \\& Hernquist 2003) with forthcoming instruments such as LOFAR ({\\it http://www.lofar.org/}) or SKA ({\\it http://www.skatelescope.org/})." }, "0403/astro-ph0403470_arXiv.txt": { "abstract": "{ Isolated HAEBE stars are believed to represent an intermediate stage of objects between young stellar objects surrounded by massive, optically thick, gaseous and dusty disks and Vega like stars surrounded by debris disks. The star \\aba \\ is already known for being surrounded by an intermediate-stage dust disk emitting a fairly large infrared and (sub-)millimetric excess. Until now, the outer disk structure has only been resolved at millimeter wavelengths and at optical wavelength coronographic imaging. We have obtained 20~$\\mu$m images which show an unexpected ellipse-shaped disk structure in emission at a distance of about 260 AU from the central star. Large azimuthal asymmetries in brightness can be noticed and the center of the ellipse does not coincide with the star. A simple, pure geometrical model based on an emission ring of uniform surface brightness, but having an intrinsic eccentricity succeeds in fitting the observations. These observations give for the first time clues on a very peculiar structure of pre-main-sequence disk geometry, i.e. a non uniform increase in the disk thickness unlike the common usual sketch of a disk with a constant flaring angle. They provide also valuable informations on the disk inclination as well as its dust composition; at such a large distance from the star, only transient heating of very small particles can explain such a bright ring of emission at mid-infrared wavelengths. Finally, the increase of thickness inferred by the model could be caused by disk instabilities; the intrinsic eccentricity of the structure might be a clue to the presence of a massive body undetected yet. ", "introduction": "Herbig Ae/Be (HAEBE) stars represent a class of intermediate mass, pre-main-sequence (PMS) stars, first described as a group by \\cite{herbig1960}. The circumstellar (CS) disks found around these stars are believed to be the sites of on-going planet formation. By studying the characteristics and evolution of the CS disk and its dust composition, valuable insights can be obtained into the processes leading to the formation of planets, and put constraints on disk and planet formation models. Infrared spectroscopy obtained with the Infrared Space Observatory (ISO) has given us insight into the dust composition of sample of isolated HAEBE systems \\citep[e.g.][]{abaur,processing,HerbigOverview}. While these spectra reveal a rich mineralogy, no direct information concerning the spatial distribution of the different dust species can be inferred from the ISO data. Most studies so far have used the available spectral energy distributions (SEDs) to put constraints on the spatial distribution of the CS material. Models for passively heated disks surrounding PMS stars are successful in reproducing the ISO spectra \\citep[e.g.][]{dullemond2001, dominik2003}, but these models cannot be uniquely constrained from SED fitting alone \\cite{degenaracy}, for this spatially resolved imaging, as presented in this paper, is required. \\begin{figure*}[t] \\vspace*{-3cm} \\parbox{11.0cm}{ \\begin{center} \\vspace*{2.5cm} \\resizebox{13.0cm}{!}{\\includegraphics[angle=90]{image.ps}} \\end{center} } \\hspace*{-0.5cm} \\parbox{6.5cm}{ \\begin{center} \\vspace*{2.5cm} \\resizebox{7cm}{10cm}{\\includegraphics[bb= 5cm 10.3cm 16cm 29.2cm,clip]{fig1_bis.ps}} \\end{center} } \\label{fig-image} \\caption{Image of the AB Aurigae disk at 20.5~$\\mu$m. The left figure shows the deconvolved image, with a pixel scale of 0.3\\arcsec/pixel. Clearly visible is a resolved central emission region surrounded by a ring like structure. The panels on the right show the normalized intensity profiles along a cut through the CAMIRAS images of 2 reference stars and AB~Aur. The upper right panel shows the comparison between two observations of reference stars, demonstrating the stability of the PSF. The panel on the lower right shows a comparison of AB~Aur with the PSF, clearly showing that the central emission is extended and that a ring-like emission structure is also detected in non deconvolved data. } \\end{figure*} Among isolated HAEBE systems, the disk around \\aba \\ is one of the most interesting and studied. Its star has a probable age of 2~Myr \\citep{mario1997}, indicating that, according to current planet formation theories, planet building could still being ongoing in this system. ISO spectra show strong PAH emission bands, and emission from silicates, and carbonaceous dust grains \\citep{mario_abaur,abaur}. Though the inferred grain sizes are differing from interstellar grains, the dust around AB~Aur seems to be relatively unprocessed, indicating an evolutionary young system. This seems also to be confirmed by combined ISO-SWS and sub-millimeter observations of H$_2$ and CO rotational lines, demonstrating that the disk still has a large gas content \\citep{Thi2001}. Its disk has been resolved in the millimeter range by \\citet{mannings1997}, showing a structure consistent with a Keplerian disk. Its surrounding nebulae and outer disk structure were also studied in the visible range using broad-band coronographic observations \\citep{grady_stis_abaur}, showing a disk with spiral-shaped structures. Near-IR interferometric \\citep{millan1999, millan2001, eisner2003}, have resolved the inner parts of the disk, showing it to be consistent with a passive disk with an inner hole, seen at a low ($\\le 40\\degr$) inclination angle. Recent mid-IR imaging \\citep{chen2003}, has also resolved the inner structure, showing it to originate from thermal emission from dust grains heated by the stellar radiation field near the central star. Here we present mid-IR imaging, not only resolving the thermal emission from dust close to the central star, but also an emission structure at the outer parts of the disk. In the next sections we will describe these observations and propose a pure geometrical model for explaining our results, unlike the paradigm according to which gravitational interactions with massive bodies are inferred when observing brightness asymmetries in dusty disks. ", "conclusions": "We have presented in this paper a thermal infrared image at 20.5 $\\mu$m of the \\aba \\ dusty disk. The deconvolved image shows an inner resolved structure containing the majority (95 \\%) of the thermal flux at this wavelength and an unexpected ring-like ellipse-shaped structure. This ring is located at an average distance of 260 AU from the star, but its most interesting features lie in large scale asymmetries and a center that is offset with respect to the star. When modeling the ring structure using a purely geometrical model (in which asymmetries are essentially produced by inclination and shadowing effects in a flared disk), we end with the conclusion that we are in presence of an {\\bf intrinsically ellipse-shaped} ring that could trace the presence of puffed-up matter from a {\\bf flared dust disk} containing transiently heated particles. Self-shadowing instabilities perturbating the disk vertical thickness could produce such a ring. However, the origin of its intrinsic eccentricity remains unclear; massive bodies could be acting in gravitationally structuring the disk. Further detailed modelings still need to be investigated to fully understand the physics of this intriguing structure; further observations, including higher resolution imaging in several PAH bands and spatially resolved mid-IR spectroscopy, are needed to better understand this object." }, "0403/astro-ph0403187_arXiv.txt": { "abstract": "We have monitored the phase of the main X-ray pulse of the Crab pulsar with the Rossi X-ray Timing Explorer (RXTE) for almost eight years, since the start of the mission in January 1996. The absolute time of RXTE's clock is sufficiently accurate to allow this phase to be compared directly with the radio profile. Our monitoring observations of the pulsar took place bi-weekly (during the periods when it was at least 30 degrees from the Sun) and we correlated the data with radio timing ephemerides derived from observations made at Jodrell Bank. We have determined the phase of the X-ray main pulse for each observation with a typical error in the individual data points of 50~$\\mu$s. The total ensemble is consistent with a phase that is constant over the monitoring period, with the X-ray pulse leading the radio pulse by 0.0102$\\pm$0.0012 period in phase, or 344$\\pm$40~$\\mu$s in time. The error estimate is dominated by a systematic error of 40~$\\mu$s in the radio data, arising from uncertainties in the variable amount of pulse delay due to interstellar scattering and instrumental calibration. The statistical error is 0.00015 period, or 5~$\\mu$s. The separation of the main pulse and interpulse appears to be unchanging at time scales of a year or less, with an average value of 0.4001$\\pm$0.0002 period. There is no apparent variation in these values with energy over the 2-30 keV range. The lag between the radio and X-ray pulses may be constant in phase (i.e., rotational in nature) or constant in time (i.e., due to a pathlength difference). We are not (yet) able to distinguish between these two interpretations. ", "introduction": "For many years it has been assumed that the main pulse and interpulse of the Crab pulsar (PSR~B0531+21) are perfectly lined up in phase over the full range of the electro-magnetic spectrum. Even though there have been reports in the past that this alignment may not be as perfect as generally assumed, the absolute calibration of spacecraft clocks was not sufficiently accurate to allow a precise measurement of the phase difference. The most compelling result predating the Rossi X-ray Timing Explorer (RXTE) observations was presented by \\citet{masn1994}, based on Figaro~II observations covering 0.15 to 4.0~MeV, that were made in 1986 and 1990. While discounting the 1986 result which has a considerable uncertainty due to potential errors in the dispersion measure, we consider the 1990 result (the $\\gamma$-ray pulse leading the radio pulse by 375$\\pm$148~$\\mu$s) fairly reliable, though the error is probably underestimated. The precision with which absolute time can be determined with the RXTE clock allows us to time X-ray pulses with an accuracy better than 10~$\\mu$s, depending on pulse shape, as shown by \\citet{rots1998b}. At the same time, the Crab pulsar monitoring program at Jodrell Bank provides timing ephemeris data, reduced to infinite frequency. These (monthly) timing ephemerides represent fits to the daily time-of-arrival measurements with rms residuals of order 20-50~$\\mu$s. This allows us to measure and monitor the radio to X-ray phase difference of the pulses with an error of about one milli-period. We have reported on these results in the past (\\citet{rots1998a}, \\citet{rots1998c}, \\citet{rots2000a}, \\citet{rots2000b}). At optical wavelengths, \\citet{sanw1999} has reported a time delay of 140~$\\mu$s (optical leading the radio), but the details are not easily accessible. \\citet{shea2003} report that, in the case of giant radio pulses, the optical pulse in the wavelength range 600-750~nm is leading the radio pulse by 100$\\pm$20~$\\mu$s. \\citet{roma2001}, on the other hand, claim that the optical (355-825~nm) and radio peaks are coincident within 30~$\\mu$s, based on test observations with a prototype transition-edge sensor detector. However, it is not clear whether the timing calibration of the instrument was complete at the time. \\citet{ulme1994} presented results from OSSE observations (50-100~keV), indicating that the hard X-ray to radio lag was $<$30$\\pm$30~$\\mu$s. It would appear that the estimate of their errors was too optimistic. \\citet{nola1993} present pulse profiles but no absolute phases. \\citet{kuip2003} report on INTEGRAL data, covering 6-50~keV, and deriving a time delay (the radio trailing) of 280$\\pm$40~$\\mu$s for a single epoch. The precise timing of the pulses in the different wavelength regimes has important repercussions for the understanding of the nature and spatial origin of the emission processes that give rise to the pulses in different parts of the spectrum. \\citet{roma1995} have suggested that, while the radio precursor comes from the polar cap, the pulse and interpulse originate in the outer gap in the magnetosphere, with the higher energy pulses being generated at significantly greater height. Thus, measuring the pulse shapes and the absolute timing throughout the electromagnetic spectrum places important constraints on the shape of the the outer gap and on the height in the magnetosphere at which the radiation is generated. In this paper we will present the results of the RXTE monitoring campaign of the Crab Pulsar from the start of that mission. We have adopted the radio nomenclature for the features in the pulse profile (main pulse, bridge, and interpulse). ", "conclusions": "The X-ray main pulse leads its radio counterpart by about 344$\\pm$40~$\\mu$s (systematic error); the statistical error is 5~$\\mu$s. This is more than twice the time difference of 140~$\\mu$s that \\citet{sanw1999} determined for the optical B band and three times the 100~$\\mu$s measured by \\citet{shea2003} in red light. The time or phase difference appears to have been constant over the past eight years, but the data are not accurate enough to distinguish between the two. The errors in individual measurements range up to 50~$\\mu$s. We should caution the reader when making comparisons with results in other wave bands. First, various authors have used different definitions of the pulse phase. In our estimation, our own definition agrees with that used in the radio band, as does the definition of \\citet{shea2003}, but that is probably not true for most other reports. Second, the scatter in values for individual observations is fairly large (a milliperiod) and may be intrinsic. Greater accuracy can only be achieved with a statistically significant set of observations. As we have mentioned, there is a systematic error of up to 40~$\\mu$s in the offset due to uncertainties in the interstellar scattering and the calibration of the radio equipment. This error may change on timescales of a few months, but since we do not know whether (and if so, by how much) the error is reduced by averaging, we quote 40~$\\mu$s as the final uncertainty for the radio to X-ray timing of the pulse. However, such an error does not affect the comparison with results from other wavebands provided they are all approximately contemporaneous and use the same Jodrell Bank timing ephemeris records. On the other hand, it also appears that most (if not all) errors for the results at other wavebands quoted in the Introduction are seriously underestimated by ignoring the systematic error. The phase difference between the two X-ray pulses is constant at 0.400 period, within the measurement errors. It is also equal to the phase difference between the radio main pulse and interpulse, within the measurement error. It may be of interest to note that in the X-ray pulse profile the trailing edges of the pulse as well as the interpulse are distinctly steeper than their leading edges. This does not appear to be the case for the optical interpulse. If the X-ray to radio lag were a true phase lag, attributable to the (radial) energy distribution across a cone, with the pulses occurring near the cone edges, one would expect the placement to be symmetrical, i.e., one X-ray pulse to be leading, the other trailing. As it stands, both X-ray pulses are leading by the same amount. The simplest explanation for this phenomenon is that we are dealing with a time delay reflecting a pathlength difference: the radio pulses originate approximately 100~km closer to the surface of the neutron star, as already suggested by \\citet{masn1994}." }, "0403/astro-ph0403378_arXiv.txt": { "abstract": "Quartessence is one of the alternatives to $\\Lambda$-CDM that has lately attracted considerable interest. According to this unifying dark matter/energy scenario, the Universe evolved from an early non-relativistic matter-dominated phase to a more recent accelerated expansion phase, driven by a single fluid component. Recently, it has been shown that some problems of the quartessence model, such as the existence of instabilities and oscillations in the matter power spectrum, can be avoided if a specific type of intrinsic entropy perturbation is considered. In the present article we explore the role of skewness in constraining this non-adiabatic scenario. We show that non-adiabatic quartessence and quintessence have different signatures for the skewness of the density distribution on large scales and suggest that this quantity might prove helpful to break possible degeneracies between them. ", "introduction": "According to the current standard cosmological model, the dynamics of the universe would be dominated by two unknown components: dark-matter (DM), responsible for structure formation, and dark-energy (DE), that causes the accelerated expansion. Although there are several candidates for both DM and DE, there is still no evidence of either of them in laboratory physics. From the point of view of simplicity, it would be interesting to explore the possibility that a single component plays the role of both DE and DM, reducing from two to one the unknown constituents of the universe. A model that provides a single description of DE and DM through \\textquotedblleft unifying-dark-matter\\textquotedblright\\ or simply quartessence \\cite{makler03} has attracted a lot of interest recently. A prototype of this model is given by the quartessence Chaplygin model (QCM) \\cite{kamenshchik01}. Both the background and linear fluctuations were extensively studied for QCM, and were compared to observational data. The generalized Chaplygin gas (as quartessence) appears to be compatible with all available data regarding the expansion history (see e.g. ref. \\cite{makler03b} and refs. therein). For adiabatic perturbations, a linear analysis was done for the CMB \\cite{amendola03}\\ and LSS \\cite{beca03}. In this case, only QCM models close to the ``$\\Lambda$CDM limit'' are allowed. Recently, it was shown that problems (pointed out in \\cite{sandvik02}), such as the existence of instabilities and oscillations in the matter power spectrum of QCM, can be avoided if a specific type of intrinsic entropy perturbation is considered. Such non-adiabatic model is consistent with the 2dF power spectrum for any value of the model parameters in the permitted interval, as long as the effective shape parameter assumes certain values \\cite{reis03b}. An \\textquotedblleft averaging problem\\textquotedblright\\ was also pointed out as a shortcoming of quartessence \\cite{avelino03}. However, it is straightforward to show that the above mentioned non-adiabatic quartessence does not suffer this kind of problem \\cite{obs}. Thus, up to the present time, we can say that adiabatic quartessence is disfavored by the data, but it is not possible to distinguish non-adiabatic quartessence from concordance models like $\\Lambda $CDM and quintessence using the previously considered observables. However, as we shall see, measurable differences in the predictions of these models appear clearly in the nonlinear regime, in particular in the skewness of the matter distribution in large scales. In our investigation we specifically consider three different quartessence models. All of them have the $\\Lambda $CDM model as a limiting case for the background solution. The analysis of these three cases indicates that our result should be applicable to more generic quartessence models. While most studies of the nonlinear regime deal with the clumping of pressureless fluid (DM), in the case of quartessence it is imperative that one includes the effects of pressure. Accordingly, we apply, and somewhat extend to include relativistic pressure, a method for the computation of density cumulants developed in refs. \\cite{bernardeau92,bernardeau94,fosalba98}. ", "conclusions": "" }, "0403/astro-ph0403652_arXiv.txt": { "abstract": "{ We present a spectroscopic analysis of two galaxy clusters at z$\\approx0.2$, out to $\\sim4$\\,Mpc. The two clusters VMF73 and VMF74 as identified by \\citet{VMFJQH98} were observed with multiple object spectroscopy using MOSCA at the Calar Alto 3.5\\,m telescope. Both clusters lie in the \\mbox{ROSAT} Position Sensitive Proportional Counter field R285 and were selected from the X-ray Dark Cluster Survey \\citep{GBCZ04} that provides optical $V$- and $I$-band data. VMF73 and VMF74 are located at respective redshifts of z$=0.25$ and z$=0.18$ with velocity dispersions of 671 km\\,s$^{-1}$ and 442 km\\,s$^{-1}$, respectively. Both cluster velocity dispersions are consistent with Gaussians. The spectroscopic observations reach out to $\\sim2.5$ virial radii. Line strength measurements of the emission lines H$_\\alpha$ and [O\\,II]$\\lambda$3727 are used to assess the star formation activity of cluster galaxies which show radial and density dependences. The mean and median of both line strength distributions as well as the fraction of star forming galaxies increase with increasing clustercentric distance and decreasing local galaxy density. Except for two galaxies with strong H$_\\alpha$ and [O\\,II] emission, all of the cluster galaxies are normal star forming or passive galaxies. Our results are consistent with other studies that show the truncation in star formation occurs far from the cluster centre. ", "introduction": "\\label{sec:Intro} Galaxy properties such as colour, morphology and spectral characteristics are functions of redshift as well as of galaxy environment. Examples for cluster specific redshift dependences are the Butcher-Oemler effect \\citep{BO78a} and the change of the ``morphological mix'' in clusters with redshift. Local clusters like the Coma cluster are dominated by S0 galaxies, which form only a small fraction in higher redshift clusters at z$\\sim0.5$, whereas spiral galaxies are more abundant in distant than in local clusters. This observation has raised the question whether the spiral galaxies in distant clusters are the progenitors of S0 galaxies in local clusters, and which kind of processes may be responsible for a transformation of one morphological type into another \\citep[e.g.][]{DOCSE97,PSDCB99}. Further evidence for an evolution of galaxy properties with redshift is provided by studies which show that the universal average star formation rate (SFR) has been rapidly declining since z$\\sim$1 \\citep[e.g.][]{LLHD96,MFDGS96,BSIK99,SPF01}. However, at a given epoch, cluster galaxies always show suppressed star formation compared with the field population at the same redshift \\citep[e.g.][]{BMYCE97}. Models of hierarchical structure formation predict a continuous accumulation of material and substructure to more and more massive galaxies, groups and clusters as time progresses \\citep[e.g.][ and references therein]{Kauff96,KCDW99b,CLBF00}. Therefore, it may be possible to link the decline in the global star formation rate to the growth of large scale structure in the universe.\\\\ A number of studies have investigated the connection between the cluster environment and morphological gradients \\citep[e.g.][]{Dress80,PG84,WG93,DOCSE97,DML01}. To characterize the cluster environment, the clustercentric distance of a galaxy as well as the local projected galaxy density are useful as independent parameters. Both of them, however, are projected parameters. The morphological dependence of cluster galaxies on the environment has been studied based on the clustercentric radius e.g. by \\citet{WG91,WG93} and based on local galaxy density by \\citet{Dress80,DOCSE97} and others. It has been discussed which parameter is more appropriate, i.e. whether galaxy morphology is influenced rather by global properties, characterized by the clustercentric distance, or by local properties like the local galaxy density. However, in centrally concentrated, regular and presumably relaxed clusters the local galaxy density is closely correlated with the clustercentric radius \\citep{DOCSE97,GNMBG03}. A correlation between morphological gradients and clustercentric distance or local galaxy density has been found, such that the dense cluster core is populated predominantly by ellipticals and S0 galaxies. Towards the cluster outskirts, the fraction of spirals increases while the elliptical and S0 fraction decreases \\citep{Dress80}.\\\\ In a similar way, the star formation activity of galaxies depends on the environment. This dependence has been investigated by various studies, which reveal a lower level of star formation in cluster galaxies compared with the field population at the same redshift \\citep{ASHCY96,BSMYC98,BBSZD02}. The results of these studies suggest that star formation is suppressed in cluster galaxies over a wide range of cluster masses. To a certain extent it is not surprising to find lower SFRs of galaxies in high density environments as these dense regions like cluster cores are dominated by galaxy types with an intrinsically lower SFR. However, recent studies show that even within the same galaxy type the SFR is lower in the cluster population compared to the surrounding field \\citep{BSMYC98,GNMBG03}.\\\\ The present analysis focuses on star formation properties of galaxies in clusters at z\\,$\\sim0.2$ out to large distances from the cluster centre. Studies of \\mbox{CNOC\\,1} data explored clusters out to $\\sim1$ virial radius (R$_v$) and found that mechanisms responsible for the low SFR of cluster galaxies are likely to start acting far out of the cluster core in the infall region, where field galaxies attracted by the cluster potential experience the influence of the cluster environment for the first time \\citep{BMYCE97,BSMYC98}. It has been shown that galaxies as far out as $\\sim2$\\,R$_{v}$ may have crossed the cluster core within a Hubble time \\citep{RS98,BNM00}. This suggests that galaxies in the cluster outskirts may have undergone significant evolution on their way through the cluster core and that cluster specific processes start acting on infalling galaxies even beyond 2\\,R$_v$. It has been discussed, for instance, whether the cluster environment induces starbursts in infalling galaxies which lead to a rapid consumption of the gas supply and a subsequent passive evolution of these galaxies \\citep{DG83,BD86,ASHCY96}. \\\\ In the local universe, the cluster environment has been explored from the densest central regions to large radial distances and into the field by studies of \\mbox{2dFGRS} and \\mbox{SDSS} data \\citep{LBDCB02,GNMBG03,BEMLB04}. At intermediate and high redshift, only few comparable studies have yet been conducted \\citep{ASHCY96,KSNOB01}. Galaxy properties, however, depend strongly on redshift. Therefore, to understand the evolution of galaxies, it is essential to investigate different environments at different redshifts. In this study we present observations of two low-mass clusters at z$\\sim0.2$ as part of a larger program to study the correlation between star formation and environment at this redshift.\\\\ The outline of the paper is as follows. In chapter 2, the selection of objects, observations and data reduction are described. The data analysis and results are presented in chapter 3. In chapter 4, the results are discussed and compared with previous studies, and chapter 5 is a summary of our findings. Throughout this paper, we use a cosmology of H$_0=70$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_m=0.3$ and $\\Omega_{\\Lambda}=0.7$ ", "conclusions": "\\label{sec:Dis} \\subsection{Comparison with earlier work} \\label{sec:comp} The relation between star formation properties of galaxies and their environment has been investigated by various studies of the local universe \\citep{LBDCB02,GNMBG03} and at intermediate redshift \\citep[e.g.][]{BBSZD02}. In all of these studies, the clustercentric distance as well as the local galaxy density are used to characterize the cluster environment.\\\\ \\citet{LBDCB02} and \\citet{GNMBG03} have analysed the star formation activity of \\mbox{2dFGRS} and \\mbox{SDSS} galaxies in the redshift range $0.05\\leq$\\,z\\,$\\leq0.1$. Both studies derive SFRs from the H$_{\\alpha}$ flux and analyse star formation properties as a function of clustercentric radius as well as of local galaxy density. \\citet{GNMBG03} investigate furthermore the environmental dependence of the H$_{\\alpha}$ and [O\\,II] emission line strengths. The trends found in the present study are qualitatively similar to the ones detected in these earlier studies. Star formation activity decreases with decreasing clustercentric radius and increasing local galaxy density. The median of both line strength distributions in the sample of VMF73 and VMF74 increases from $\\sim0$\\,\\AA\\ in the centre to $\\sim5$\\,\\AA\\ in the outer regions. This is in good agreement with \\citet{GNMBG03}. The same study finds density dependences of the star formation activity such that the median of the H$_\\alpha$ and [O\\,II] line strength distributions decreases with increasing local galaxy density. The density estimates used by \\citet{GNMBG03} are different from the ones in the present study. \\citet{GNMBG03} apply a background subtraction using redshifts while in the present analysis a statistical background subtraction was applied. As a result, the densities in Figure~\\ref{fig:densEm} are much larger than those of \\citet{GNMBG03} due to the difference in projection. In the VMF73 and VMF74 clusters, we see a qualitatively similar effect in H$_\\alpha$ and [O\\,II]; however, the sample size is too small to warrant a detailed comparison.\\\\ In a redshift range comparable to the one of the present analysis, \\citet{BBSZD02} have carried out a study of ten galaxy clusters with low X-ray luminosity. Galaxy clusters of low X-ray luminosity have been chosen in order to study the environmental dependence of galaxy SFRs and possible mechanisms suppressing the star formation in low mass clusters which are considered the progenitors of more massive clusters in models of hierarchical structure formation. The cluster sample of VMF73 and VMF74 shows an increase in the mean and median of the [O\\,II] line strength distribution from the centre to the outer regions which is consistent with \\citet{BBSZD02}. However, there is large scatter in the data of VMF73 and VMF74 and the observed trends in mean and median are of limited statistical significance. A stronger trend is found in the increase of the galaxy fraction with W$_0$([O\\,II])\\,$>5$\\,\\AA\\ from $\\sim15$\\% in the centre to $\\sim50$\\% in the outer fields. \\citet{BBSZD02} find an increase in the fraction of star forming galaxies from $\\sim10$\\% in the centre to $\\sim30$\\% at $\\sim1$\\,R$_v$. Density trends in the sample of VMF73 and VMF74 have been detected such that the mean and median of the [O\\,II] line strength distribution decrease with increasing density, while \\citet{BBSZD02} find that the mean reaches $\\sim6$\\,\\AA\\ only in the lowest density regions. Both studies detect similar density trends in the fraction of galaxies with W$_0$([O\\,II])\\,$>5$\\,\\AA\\@.\\\\ As a further indicator for star formation, galaxy colours in the two clusters VMF73 and VMF74 were investigated. We find a radial trend in the colours such that with increasing radius the average colours become bluer. At large radii beyond $\\sim1.5$\\,R$_v$, there are few galaxies with a dominant old, red population. On the other hand, there are some rather blue galaxies without emission found in the central regions. This population may consist of galaxies which had ongoing star formation at the time of their infall into the cluster environment. On their way through the cluster, the diffuse gas in the halo may have been lost which therefore cannot replenish the fuel for star formation in the disk and the SFR slowly declines \\citep{BNM00}. However, this mechanism predicts a systematically lower SFR of star forming galaxies in high density environments compared with similar galaxies in lower densities. A recent study of SDSS and 2dF data finds that the lower level of star formation in high density environments is largely due to the smaller fraction of star forming galaxies in these environments, while there is little difference in the star forming galaxy populations in high and low density regimes \\citep{BEMLB04}. The sample of VMF73 and VMF74 is too small to draw any strong conclusions from the population of blue galaxies without H$_\\alpha$ emission. Figure~\\ref{fig:vicHab} suggests that there is a small difference in the H$_\\alpha$ emission between blue, star forming galaxies in the inner and outer fields. Furthermore, in the VMF73 and VMF74 sample a distinctive colour change is visible at a clustercentric radius of $\\sim1$\\,R$_v$ where the red sequence of both clusters disappears. This feature is reminiscent of the work of \\citet{KSNOB01} who find an abrupt colour change at a characteristic density of 2 galaxies h$^2_{50}$ Mpc$^{-2}$ in the cluster Cl~0939+4713 at z$=0.41$. \\subsection{Physical Implications} The aim of this investigation has been to explore the star formation properties of galaxies over a wide range of densities with a special focus on galaxies in large distances from the cluster core. The difference between the present study and the two previous ones cited above is that we investigated cluster galaxies in a higher redshift regime than did \\citet{GNMBG03} analysing \\mbox{SDSS} data. Furthermore, this work reaches out to larger clustercentric distances than the study of X-ray faint clusters at intermediate redshift \\citep{BBSZD02}. In regions as far out as $\\sim3-4$\\,R$_v$ from the cluster centre, galaxies from the surrounding field are thought to fall continuously into the cluster attracted by its gravitational potential and experience for the first time the influence of the cluster environment. Therefore, galaxies in these infall regions which denote a transition between the low density field and the cluster are expected to have intermediate properties between the field and cluster galaxy population. The results of \\citet{GNMBG03} and \\citet{BBSZD02} are consistent with this picture in finding an increase in line strengths and SFRs with increasing clustercentric distance and decreasing density. The present study finds similar radial and density dependences of galaxy star formation. The change of galaxy colours with radius is a further indication for some kind of transformation of cluster galaxies. The observed disappearance of the red sequence at a radius of $\\sim1$\\,R$_v$ is similar to an abrupt colour break reported by \\citet{KSNOB01} at a characteristic density of 2 galaxies h$^2_{50}$ Mpc$^{-2}$. \\citet{KSNOB01} find that this density corresponds to subclumps outside the cluster centre. This may be a further hint that galaxy transformation first occurs in groups and subclumps far from the centre.\\\\ Several processes are being discussed in the literature as candidates for suppressing star formation in cluster galaxies. The radial dependence of star formation properties suggests that this mechanism starts to act about 2.5\\,R$_v$ from the cluster core and possibly even beyond. This result rules out mechanisms that are effective only in the highest density regimes, such as ram pressure stripping, as being solely responsible for the decreasing SFRs throughout the cluster. Although more than one process may be involved and ram pressure stripping may still play a role in the cluster core, it is likely that in the outer regions a more subtle mechanism is at work which does not halt the star formation abruptly. Apparently, this mechanism starts to affect galaxies relatively early after their first contact with the cluster environment, for even galaxies far out of the central regions show a reduced star formation. The scenario of strangulation \\citep{BNM00} in which the diffuse gas in the dark halo of an infalling galaxy is lost may be an explanation for the reduced SFRs in cluster galaxies. As discussed above, this mechanism predicts lower SFRs of blue galaxies in high density regions compared with blue galaxies in lower density regimes. However, based on the blue galaxy population in the VMF73 and VMF74 cluster sample alone, we cannot draw any strong conclusions about the mechanism suppressing star formation in the cluster environment. The observed decrease in SFRs is compatible with strangulation, although it is possible that other effects such as galaxy-galaxy interactions play a role or were important in the past and influence galaxy evolution in small groups even before a galaxy enters the cluster environment. \\subsection{Outlook} The reduction and analysis of the remaining four galaxy clusters will increase the sample size and enable us to investigate the environmental dependence of galaxy star formation properties with more precision and statistical significance. With the complete sample of six galaxy clusters, both the SFR-radius relation and the SFR-density relation can be studied without the limitations of low number statistics and thus help us to understand the evolution of galaxies in the cluster environment.\\\\ Additionally, it would be interesting to study the dependence of star formation activity on the X-ray luminosity of clusters. Groups and low mass X-ray faint clusters are considered the progenitors of more massive clusters in hierarchical merging models. Studying clusters over a range of masses is therefore an important step to investigate the growth of structure in the universe." }, "0403/astro-ph0403008_arXiv.txt": { "abstract": "s{We model the spectral energy distribution of the ultrasoft broad-line AGN RE J2248-511 with Comptonised accretion disc models. These are able to reproduce the steep optical and ultrasoft X-ray slopes, and the derived black hole mass is consistent with independent mass estimates. This AGN displays properties of both broad and narrow line Seyfert 1 galaxies, but we conclude that it is intrinsically a `normal' Seyfert 1 viewed at high inclination angle.} ", "introduction": "RE J2248-511 is a nearby ($z=0.101$) EUV-selected Seyfert galaxy discovered by the \\emph{ROSAT} Wide Field Camera\\cite{Po}. Further observations showed this source to have a strong soft X-ray excess and spectral variability at optical\\cite{Ma} \\cite{Gru} and soft X-ray\\cite{emp95} \\cite{aad} wavelengths. However, none of these observations were simultaneous, so the existence of an optical to soft X-ray big blue bump (BBB) could not be confirmed. The soft X-ray spectrum resembles those of narrow-line Seyfert 1 galaxies (NLS1's), but this AGN has high velocity optical emission lines\\cite{Ma} with H$\\beta$ FWHM$\\sim$2900 km s$^{-1}$, which classifies it as a Seyfert 1. Studies of \\emph{ROSAT} PSPC slopes in AGN had concluded that sources with both steep soft X-ray continuum slopes and broad optical emission lines are not found in nature\\cite{BBF} \\cite{WB} \\cite{Gru}. This makes RE J2248-511 an unusual and ideal case in which to examine the relationship between the X-ray and optical continua, specifically the interpretation of the BBB as thermal emission from an accretion disc. We describe the results of a multiwavelength monitoring campaign of RE J2248-511, consisting of X-ray observations from the \\emph{XMM} satellite with supporting quasi-simultaneous optical observations made at the South African Astronomical Observatory (SAAO) and archival multiwavelength data. ", "conclusions": "We have shown that Comptonised accretion disc models, which treat the optical to soft X-ray emission as a single BBB, are able to comprise the majority of this flux. From these we derive constraints on the black hole mass which are fully consistent with the mass we obtain through the independent photoionisation technique. This implies that thermal disc emission is the likely origin of the optical, UV and some of the soft X-ray continuum. An accretion disc does not, however, constitute the soft excess as observed with the pn, but it could be the origin of the `ultrasoft' component below 0.25 keV observed with \\emph{ROSAT}. The 0.3-2 keV soft component observed with \\emph{XMM} has a blackbody-like shape, but is too hot to be blackbody disc emission. A model including Comptonisation of soft photons in a hot plasma provides a good fit. Between the \\emph{XMM} observations, the X-ray spectral shape of RE J2248-511 remained approximately constant, while comparison with previous X-ray data shows long-term variability. This demonstrates that the `soft' state is a long-lived phase, and not, for example, a rapid flaring of the disc. Is RE J2248-511 an intermediate class of object linking the Seyfert 1's with the NLS1's, or a true Seyfert 1 seen with a particular observational bias? \\\\ Contrary to the proposed NLS1 scenario\\cite{PDO}, the data are best fitted with high black hole masses ($\\sim10^{8}$ M$_{\\odot}$), whilst still favouring the high accretion rates suggested for NLS1 galaxies. Therefore, it is not necessary for a black hole to have a low mass for the formation of an ultrasoft X-ray excess. Comparison of Comptonised accretion discs to the SED show that the orientation of the disc is close to face-on, allowing us to see a greater surface area of the accretion disc. If most of the EUV emission arises in the accretion disc then this source would appear EUV-bright compared with similar Seyfert galaxies viewed more edge-on. In the hard X-rays RE J2248-511 resembles a normal Seyfert 1 far more than a NLS1\\cite{BME}. The soft X-ray flux measured with \\emph{XMM} is typical of a Seyfert 1\\cite{TP}, and its slope is similar to that found in the PG bright quasar sample \\cite{Laor}. This source also follows the observed correlation between Balmer linewidth and soft X-ray slope\\cite{emp92} \\cite{Laor} if the `ultrasoft' part of the X-ray spectrum (below 0.25 keV) is excluded. Variability in the soft X-ray excess is a property which RE J2248-511 shares with the NLS1 galaxies, but changes are often more dramatic in NLS1's than observed here. The strength and broad wavelength span of the BBB in RE J2248-511 is unusual for a broad-line AGN, but we find this may be explained by a face-on Comptonised accretion disc, likely to be accreting at a high rate onto a 10$^{7.5}$-10$^{8.5}$ M$_{\\odot}$ black hole. We propose that this source is intrinsically a normal Seyfert 1, but shares the ultrasoft X-ray excess property with the NLS1's because it is observed at a higher inclination angle than the majority of Seyfert 1 galaxies." }, "0403/astro-ph0403522_arXiv.txt": { "abstract": "Here we report our discovery of a band of blue luminescence (BL) in the Red Rectangle (RR) nebula. This enigmatic proto-planetary nebula is also one of the brightest known sources of extended red emission as well as of unidentified infra-red (UIR) band emissions. The spectrum of this newly discovered BL is most likely fluorescence from small neutral polycyclic aromatic hydrocarbon (PAH) molecules. PAH molecules are thought to be widely present in many interstellar and circumstellar environments in our galaxy as well as in other galaxies, and are considered likely carriers of the UIR-band emission. However, no specific PAH molecule has yet been identified in a source outside the solar system, as the set of mid-infra-red emission features attributed to these molecules between the wavelengths of 3.3 $\\mu$m and 16.4 $\\mu$m is largely insensitive to molecular sizes. In contrast, near-UV/blue fluorescence of PAHs is more specific as to size, structure, and charge state of a PAH molecule. If the carriers of this near-UV/blue fluorescence are PAHs, they are most likely neutral PAH molecules consisting of 3-4 aromatic rings such as anthracene (C$_{14}$H$_{10}$) and pyrene (C$_{16}$H$_{10}$). These small PAHs would then be the largest molecules specifically identified in the interstellar medium. ", "introduction": "The family of emission bands at 3.3, 6.2, 7.7, 8.6, 11.2, \\& 12.7 $\\mu$m (called the unidentified infra-red (UIR) bands) is found in almost all astrophysical environments including the diffuse interstellar medium (ISM), the edges of molecular clouds, reflection nebulae, young stellar objects, HII regions, star forming regions, some C-rich Wolf-Rayet stars, post-AGB stars, planetary nebulae, novae, normal galaxies, starburst galaxies, most ultra-luminous infra-red galaxies and AGNs \\citep[see][and references therein]{pet04}. Approximately 20-30\\% of the Galactic IR radiation is emitted in these UIR bands and 10-15\\% of the interstellar carbon is contained in the UIR carriers \\citep{sw95}, indicating that the carriers represent an abundant component of the ISM. These UIR bands are the signatures of aromatic C-C and C-H fundamental vibrational and bending modes, and are generally attributed to a family of PAH molecules containing 50-100 carbon atoms \\citep{hony01,cook98,lp84,ver01,atb85,sel84}. Although the presence of PAHs in space is widely accepted by most astronomers, their specific sizes and ionization states remain elusive. On absorption of a far-UV photon, a PAH molecule usually undergoes a transition to an upper electronic state. If the molecule undergoes iso-energetic transitions to highly vibrationally excited levels of the ground state, then the molecule relaxes through a series of IR emissions in the C-C and C-H vibrational and bending modes. These transitions are largely independent of size, structure and ionization state of the molecule. However, \\emph{electronic} fluorescence, a transition from the upper excited level to the ground state, is more specific \\citep{reyle00}. In particular, the wavelength of the first electronic transition seen in neutral PAHs is closely dependent on the size of the molecular species. In general, this fluorescence wavelength increases with the molecular weight of the molecule (Figure 1). An observation of fluorescence in an astronomical source in the UV/visible range offers the possibility of estimating the size of the PAH molecules, which the observation of the UIR band emission in the same source does not. \\begin{figure} \\plotone{f1.eps} \\caption{The first electronic transition in the fluorescence spectra \\citep{berl65,peaden80,reidel1,reidel2} of neutral PAH molecules as a function of their molecular weight. We distinguish between condensed and non-condensed PAHs, with condensed PAHs considered to be more stable under interstellar conditions.\\label{fig1}} \\end{figure} ", "conclusions": "Another important factor which should be considered when identifying possible carriers is fluorescence efficiency. Although many different fluorescing species may exist with similar abundances, the spectra of the most efficient ones will be dominant. Laboratory studies have shown that on comparison of emission rates per molecule, from PAH molecules placed at 1 AU from the Sun, anthracene and pyrene show 10-100 times more efficiency than naphthalene (C$_{10}$H$_8$) and phenanthrene (C$_{14}$H$_{10}$) \\citep{brech94}. Additional support for small PAHs as sources of the observed BL is found in the spatial correlation between the BL and the PAH 3.3 $\\mu$m emission in the RR and the presence of the distinct PAH ionization discontinuity in the spectrum of the central source, HD44179 (U. Vijh. A. Witt, \\& K. Gordon, in preparation)." }, "0403/astro-ph0403714_arXiv.txt": { "abstract": "The observed number counts of quasars may be explained either by long-lived activity within rare massive hosts, or by short-lived activity within smaller, more common hosts. It has been argued that quasar lifetimes may therefore be inferred from their clustering length, which determines the typical mass of the quasar host. Here we point out that the relationship between the mass of the black-hole and the circular velocity of its host dark-matter halo is more fundamental to the determination of the clustering length. In particular, the clustering length observed in the 2dF quasar redshift survey is consistent with the galactic halo -- black-hole relation observed in local galaxies, provided that quasars shine at $\\sim 10$--$100\\%$ of their Eddington luminosity. The slow evolution of the clustering length with redshift inferred in the 2dF quasar survey favors a black-hole mass whose redshift-independent scaling is with halo circular velocity, rather than halo mass. These results are independent from observations of the number counts of bright quasars which may be used to determine the quasar lifetime and its dependence on redshift. We show that if quasar activity results from galaxy mergers, then the number counts of quasars imply an episodic quasar lifetime that is set by the dynamical time of the host galaxy rather than by the Salpeter time. Our results imply that as the redshift increases, the central black-holes comprise a larger fraction of their host galaxy mass and the quasar lifetime gets shorter. ", "introduction": "The Sloan Digital Sky Survey (York et al.~2000) and the 2dF quasar redshift survey (Croom et al.~2001a) have measured redshifts for large samples of quasars, and determined their luminosity function over a wide range of redshifts (Boyle et al.~2000; Fan et al.~2001a,b; Fan et al.~2003). The 2dF survey has also been used to constrain the clustering properties of quasars (Croom et al. 2001b). It has been suggested that quasars have clustering statistics similar to optically selected galaxies in the local universe, with a clustering length $R_0\\sim8$Mpc. The large sample size of the 2dF quasars also provided clues about the variation of clustering length with redshift (Croom et al. 2001b) and apparent magnitude (Croom et al.~2002). Quasars appear more clustered at high redshift, although with a relatively mild trend. There is also evidence that more luminous quasars may be more highly clustered. As pointed out by Martini \\& Weinberg~(2001) and Haiman \\& Hui~(2001), the quasar correlation length determines the typical mass of the dark matter halo in which the quasar resides. One may therefore derive the quasar duty-cycle by comparing the number density of quasars with the density of host dark matter halos. The quasar lifetime then follows from the product of the duty-cycle and the lifetime of the dark-matter halo in between major mergers, although there is a degeneracy between the lifetime and the quasar occupation fraction or beaming. Preliminary results suggested quasar lifetimes of $t_{\\rm q}\\sim10^6-10^7$ years, consistent with the values determined by other methods (see Martini~2003 for a review), including the transverse proximity effect (Jakobsen et al.~2003) and counting arguments relative to the local population of remnant supermassive black holes (SMBHs; see Yu \\& Tremaine~2002). Kauffmann \\& Haehnelt~(2002) used a detailed semi-analytic model (Kauffmann \\& Haehnelt~2000) to predict the correlation length of quasars and its evolution with redshift. They found that their model reproduces the correct correlation length as well as its redshift evolution for present-day quasar lifetimes of $t_{\\rm q}\\sim10^7$ years. In this work we argue that the correlation length of quasars is fundamentally determined by the relation between the masses of SMBHs and their host galactic halo rather than by the quasar lifetime. We find that the amplitude of the correlation length depends on the product of the SMBH--halo relation and the typical fraction of the Eddington luminosity at which quasars shine. Moreover, we show that the evolution of the correlation length is sensitive to how the SMBH -- halo relation evolves with redshift, and therefore to the physics of SMBH formation and quasar evolution. The paper is organized as follows. In \\S~\\ref{cf} and \\S~\\ref{MbhMhalo} we discuss the calculation of the correlation function of quasars with a particular apparent magnitude $B$, and the different scenarios for the evolution of the SMBH -- galaxy halo relation. In \\S~\\ref{cfobs} we compare fiducial model correlation functions to the results of the 2dF quasar redshift survey (Croom et al.~2001b,c). We find that the shallow evolution of the correlation length implies that SMBH mass has a redshift-independent scaling with the circular velocity of the host dark matter halo rather than with its mass. The ranges of the normalizations in the SMBH -- galaxy halo relation and of the fraction of Eddington allowed by observations of the observed quasar correlation function are explored in \\S~\\ref{epseta}. In \\S~\\ref{lf} we examine the quasar lifetime by requiring that observational constraints from both the correlation and luminosity functions be satisfied simultaneously. Finally, in \\S~\\ref{disc} we summarize our results and discuss a very simple, physically motivated model that satisfies all constraints with no free parameters. Throughout the paper we adopt the set of cosmological parameters determined by the {\\em Wilkinson Microwave Anisotropy Probe} (WMAP, Spergel et al. 2003), namely mass density parameters of $\\Omega_{m}=0.27$ in matter, $\\Omega_{b}=0.044$ in baryons, $\\Omega_\\Lambda=0.73$ in a cosmological constant, and a Hubble constant of $H_0=71~{\\rm km\\,s^{-1}\\,Mpc^{-1}}$. ", "conclusions": "\\label{disc} In this paper we have demonstrated that the relationship between the mass of a SMBH and its host dark-matter halo is {\\it the} fundamental quantity that determines the clustering of quasars. In particular the $M_{\\rm bh}$--$M_{\\rm halo}$ relation and its evolution with redshift, may be determined directly from the evolution in the correlation length of quasars, independent of consideration of the quasar luminosity function or assumptions about the quasar lifetime. Beginning with the locally observed $M_{\\rm bh}$--$M_{\\rm halo}$ relation, we have shown that the observed correlation length of quasars in the 2dF quasar redshift survey is consistent with SMBHs that shine at $\\sim 10$--$100\\%$ of their Eddington luminosity during their bright quasar episode. Moreover, the evolution of the clustering length with redshift is consistent with a SMBH mass whose redshift-independent scaling is with the circular velocity of the host dark matter halo (see equation~\\ref{eps}). In contrast, it appears that a relation for the SMBH mass whose redshift-independent scaling is with the mass of the host dark matter halo (equation~\\ref{eps2}) would have resulted in too much evolution of the clustering length with redshift. The portion of the 2dF quasar survey from which the data used in this study were drawn, contains $\\sim10^4$ quasars. Upon completion, the Sloan Digital Sky Survey will have spectra for ten times this number of quasars, spread over a larger redshift range (York et al.~2000). This will allow more accurate determination of the correlation length and its evolution, and hence provide more accurate constraints the SMBH population. The selection of quasars for the spectroscopic survey of SDSS at $z<3$ is restricted to $i^*<19$. Thus, most of the low redshift SDSS quasars will fall at luminosities that are brighter than the characteristic break in the luminosity function (e.g. Boyle et al.~2000). These bright quasars have a lower abundance than expected from the number of halos merging at $z\\la 2$. The reason for this low abundance may be either that the lifetime (or occupation fraction) of bright quasars is lower than expected at $z\\la 2$, or alternatively that the fraction of the Eddington limit at which the quasars shine is small. The 2dF quasar sample straddles the luminosity function break. We have shown that the similarity in the clustering statistics of sub-samples of quasars having different ranges in luminosity suggest that the lifetime (or occupation fraction) of bright quasars is lower than expected at $z\\la 2$, but that these quasars also shine near their Eddington luminosity. Measurements of the correlation length as a function of quasar luminosity in SDSS can help further distinguish between these two possibilities. As we have shown in a previous paper (Wyithe \\& Loeb~2003), feedback--regulated growth of SMBHs during active quasar phases implies $M_{\\rm bh}\\propto v_c^5$, consistent with the inferred relation for galactic halos in the local universe (Ferrarese 2002). Assuming $\\eta=1$ as suggested by observations of both low and high-redshift quasars (Floyd~2003; Willott, McLure \\& Jarvis~2003), we have shown that only $\\sim7\\%$ of the Eddington luminosity output needs to be deposited into the surrounding gas in order to unbind it from the host galaxy over the dynamical time of the surrounding galactic disk. The power-law index of 5 in the $M_{\\rm bh}$--$v_c$ relation, is larger than the values of 4--4.5 inferred from the local relation between $M_{\\rm bh}$ and the stellar velocity dispersion $\\sigma_\\star$ (Merritt \\& Ferrarese 2001; Tremaine et al. 2002); the difference originating from the observation that the $v_c$--$\\sigma_\\star$ relation is shallower than linear for stellar bulges embedded in cold dark matter halos (Ferrarese~2002). Feedback--regulated growth also implies that the $M_{\\rm bh}$--$v_{\\rm c}$ relation should be independent of redshift, as implied by the slow evolution in the clustering length of quasars. A very simple model can therefore be constructed to describe the correlation length of quasars and its evolution in the 2dF survey, as well as the number counts of quasars out to very high redshift ($z\\sim 6$). The model includes four simple assumptions, but no free parameters: {\\it (i)} the locally observed $M_{\\rm bh}$--$M_{\\rm halo}$ relation extends to high redshifts through equation~(\\ref{eps}) with the SMBH mass scaling as halo circular velocity to the 5-th power; {\\it (ii)} SMBHs shine near their Eddington luminosity with a universal spectrum (Elvis et al.~1994) during luminous quasar episodes; (iii) quasar episodes are associated with major galaxy mergers; and (iv) the quasar lifetime is set by the dynamical time of the host galaxy. The assumption of a SMBH mass that scales with halo circular velocity independently of redshift is supported by the observations of Shields et al.~(2003) that there is no evolution in the $M_{\\rm bh}$--$\\sigma_\\star$ relation out to $z\\sim3$. The assumption that quasars shine near their Eddington limit is supported by observations of high and low redshift quasars (Floyd~2003; Willott, McLure \\& Jarvis~2003). Overall, the sample of available data on the clustering properties and the number counts of quasars is most readily explained by a quasar lifetime that is set by the dynamical time of the host galaxy rather than by the Salpeter (1964) $e$-folding time for the growth of its mass." }, "0403/astro-ph0403591_arXiv.txt": { "abstract": "{ A sample of 229 nearby thick disk stars has been used to investigate the existence of an age-metallicity relation (AMR) in the Galactic thick disk. The results indicate that that there is indeed an age-metallicity relation present in the thick disk. By dividing the stellar sample into sub-groups, separated by 0.1\\,dex in metallicity, we show that the median age decreases by about 5--7\\,Gyr when going from [Fe/H]\\,$\\approx -0.8$ to [Fe/H]\\,$\\approx -0.1$. Combining our results with our newly published $\\alpha$-element trends for a local sample of thick disk stars, that show signatures from supernovae type Ia (SN\\,Ia), we can here draw the conclusion that the time-scale for the peak of the SN\\,Ia rate is of the order 3--4 Gyr in the thick disk. The tentative evidence for a thick disk AMR that we present here also has implications for the thick disk formation scenario; star-formation must have been an ongoing process for several billion years. This is further discussed here and appear to strengthen the hypothesis that the thick disk originates from a merger event with a companion galaxy that puffed up a pre-existing thin disk. ", "introduction": "Chemical evolution of stellar populations is an important part of any model of galaxy formation and evolution. Many studies in the past decades show how we are able to further refine and constrain models of Galactic chemical evolution by combining kinematics and elemental abundances of local dwarf stars (e.g. Chiappini et al.~\\cite{chiappini}; Matteucci~\\cite{matteucci}; Edvardsson et al.~\\cite{edvardsson}; Feltzing \\& Gustafsson~\\cite{feltzing1998}; Feltzing et al.~\\cite{feltzing2003}; Bensby et al.~\\cite{bensby}; Reddy et al.~\\cite{reddy}). As evidenced by the cited articles our understanding of chemical evolution is mainly driven by the studies in the solar neighbourhood but have far reaching impact for our interpretation of integrated light from other galaxies (e.g. Matteucci~\\cite{matteucci}). However, it is not only the elemental abundances and kinematics of the stars that are of importance when we want to further improve the models of galaxy formation and evolution but also the ages of the stars (see e.g. Edvardsson et al.~\\cite{edvardsson}; Raiteri et al.~\\cite{raiteri}; Pilyugin \\& Edmunds~\\cite{pilyugin}; Berczik~\\cite{berczik}). Many studies have found there to be a clear relation between the ages and the metallicities of the solar neighbourhood disk stars (Twarog~\\cite{twaroga}, \\cite{twarogb}; Rocha-Pinto et al.~\\cite{rochapinto}; Meusinger et al.~\\cite{meusinger}). In contrast to this Edvardsson et al.~(\\cite{edvardsson}) found no particular evidence for an age-metallicity relation in the Galactic disk in the solar neighbourhood and Feltzing et al.~(\\cite{feltzing}) confirmed this. Feltzing et al.~(\\cite{feltzing}) also showed how dangerous selection effects could be and how an artificial age-metallicity relation can be created (see their Figs.~13 and 14). Gilmore \\& Reid~(\\cite{gilmore}) showed that our galaxy is host to two kinematically distinct disk structures. The ``new'' disk was dubbed the thick disk and was found to have a mean metallicity around $-0.6$\\,dex (Wyse \\& Gilmore~\\cite{wyse}) and a scale-height of 800--1300\\,pc (e.g. Reyl\\'e \\& Robin~\\cite{reyle}; Chen~\\cite{chen2}) while the thin disk has a mean metallicity of around $-0.1$\\,dex and a scale height of 100--300 pc (e.g. Gilmore \\& Reid~\\cite{gilmore}; Robin et al.~\\cite{robin}). Recent studies have shown that stars selected to belong to either the thin or the thick disk show different trends for the elemental abundances (e.g. Fuhrmann~\\cite{fuhrmann}; Feltzing et al.~\\cite{feltzing2003}; Bensby et al.~\\cite{bensby}; Bensby et al.~\\cite{bensby_syre}; Reddy et al.~\\cite{reddy}; Prochaska et al.~\\cite{prochaska}; Mashonkina \\& Gehren~\\cite{mashonkina}). The question then arises: could it be so that the lack of a relation between ages and metallicities for stars in the solar neighbourhood is in fact a population effect? That is, are we looking at a mixture of stars from (at least) two populations that have different chemical enrichment histories? It thus appears natural to, yet again, revisit the question of an age-metallicity relation in the solar neighbourhood. In the study presented here we will address the question of a relation between ages and metallicities for stars that are kinematically selected to resemble the thick disk closely. The paper is organized as follows: in Sect.~\\ref{sect:kin} we describe the stellar sample and the kinematical selection criteria and investigates if there are potential biases present in the sample. In Sect.~\\ref{sect:alpha} we describe the choice of $\\alpha$-enhancement used in the isochrones when deriving the stellar ages. In Sect.~\\ref{sec:amr} we derive ages from stellar isochrone fitting and find that there is a possible age-metallicity relation present in the thick disk. In Sect.~\\ref{sec:discussion} we discuss the implications this tentative age-metallicity relation have on the star-formation history of the thick disk, on the time-scale of SN\\,Ia rate in the thick disk, and on our understanding of the origin and evolution of the thick disk. Finally, in Sect.~\\ref{sec:summary} we give a short summary. ", "conclusions": " \\begin{enumerate} \\item \tThere is an age-metallicity relation present in the Galactic thick disk, indicating that it has had an ongoing star formation for a time-period of up to 5\\,Gyr (this is model dependent). \\item\tThe thick disk age-metallicity relation in combination with the abundance trends for $\\alpha$-elements in the thick disk, that show signatures from SN\\,Ia, indicate that the time-scale for the peak of the SN\\,Ia rate in the thick disk is of the order 3--4\\,Gyr. \\item\tThe quite long star formation period in the thick disk strengthens the hypothesis that the thick disk formed as a result of an ancient merger event between the Milky Way and a companion galaxy. \\end{enumerate} Studies of the thick disk using nearby stars will always be be subject to uncertainties due to the overlapping velocity-, and metallicity distributions of the thin and thick disk. At distances well above $Z\\approx1.5$\\,kpc from the Galactic plane the thick disk is the dominant stellar population. Determining accurate metallicities and $\\alpha$-abundances for a larger sample of dwarf stars at high $Z$ would therefore enable accurate age determinations that would verify the existence (or non-existence) of the thick disk AMR deduced from nearby stars." }, "0403/astro-ph0403558_arXiv.txt": { "abstract": "It has been conjectured that the distribution of magnifications of a point source microlensed by a randomly distributed population of intervening point masses is independent of its mass spectrum. We present {\\it gedanken} experiments that cast doubt on this conjecture and numerical simulations that show it to be false. ", "introduction": "Every investigation of microlensing at high optical depth that has explored the effect of multiple microlens mass components has led to the conclusion that the magnification probability distribution is independent of the spectrum of microlens masses. The recent effort by \\citet{2001MNRAS.320...21W} is typical. While it was not their principal result, they comment in passing \\begin{quotation} ``... we confirm the finding of \\citet{1992ApJ...386...19W} and \\citet{1995MNRAS.276..103L} that the magnification distribution is independent of the mass function.'' \\end{quotation} This conjecture has important consequences regarding the more general applicability of microlensing studies that are limited to a single mass component. While galaxies have stars with a range of masses, restricting to a single component makes analytic calculations more tractable \\citep[e.g.][]{ 1986MNRAS.223..113P,1987ApJ...319....9S,1997ApJ...489..508K} and greatly decreases the number of cases that must be simulated numerically \\citep[e.g.][]{1992ApJ...386...19W,1995MNRAS.276..103L,2001MNRAS.320...21W}. If true, the conjecture simplifies things considerably. Both theoretical and experimental lines of evidence lead to this conclusion, which has struck many investigators as obvious. On the experimental side, simulations like those carried out by \\citet{2001MNRAS.320...21W} and their predecessors produce magnification histograms for different mass distributions that appear to be indistinguishable for fixed surface mass density and shear. On the theoretical side, the high magnification tail of the magnification probability distribution has been shown to be independent of the microlens mass spectrum \\citep{1987ApJ...319....9S}. Moreover, \\citet{1992A&A...258..591W} showed that the average number of positive parity microimages depends only upon the surface mass density (or equivalently the convergence) and the shear. Since the scale free nature of gravity requires that the magnification probability distribution for a point source be the same for microlensing by a single mass of any size, it would appear strange if a mixture of two masses (at constant convergence and shear) produced a different magnification probability distribution. There is, however, at least one argument against this apparently obvious conclusion, which we detail in $\\S$ 2 below. It suggests that the magnification probability distribution {\\it does} depend upon the mass spectrum. The argument suggests that the dependence would show up in a highly magnified negative parity macroimage -- typically one of a close pair of images in a quadruply imaged quasar like PG1115+080. We have carried out lensing simulations of such an image (at constant convergence and shear) for a variety of different cases. In Figure 1 we show simulations with two populations of point masses. The first component is comprised of $1.000 \\msun$ objects referred to hereafter as ``micro-lenses.'' The second component is comprised of $0.005 \\msun$ objects referred to hereafter as ``nano-lenses.'' The designations and mass scale are arbitrary but are intended to convey the sense that the micro-lenses are very much smaller than the lensing galaxy and that the nano-lenses are very much smaller than the micro-lenses. The eight panels of Figure 1 show magnification histograms obtained by varying the mass fractions in the micro-lensing component, with the remaining fraction in the nano-lensing component. For the sake of comparison, we reproduce in each panel the result for a pure micro-lensing component. As the fraction contributed by micro-lenses decreases to 20\\% and 10\\% the histogram broadens out and develops a second peak. But as it decreases further to 0\\%, the magnification distribution narrows and ends up looking like the 100\\% case (modulo finite source effects and sample variance). Unless our simulations are faulty, the conjecture is false. In $\\S$ 2 we put forward a qualitative argument for the dependence of the microlensing probability distribution on the mass spectrum. In $\\S$ 3 we give details of the numerical simulations that confirm the effect. In $\\S$ 4 we offer a qualitative interpretation of our results. In $\\S$ 5 we discuss some astrophysical consequences. ", "conclusions": "" }, "0403/astro-ph0403244_arXiv.txt": { "abstract": "Young star clusters with masses similar to those of classical old globular clusters are observed not only in starbursts, mergers or otherwise disturbed galaxies, but also in normal spiral galaxies. Some young clusters with masses as high as $\\sim10^6 \\, \\msun$ have been found in the disks of isolated spirals. Dynamical mass estimates are available for a few of these clusters and are consistent with Kroupa-type IMFs. The luminosity (and possibly mass-) functions of young clusters are usually well approximated by power-laws. Thus, massive clusters at the tail of the distribution are naturally rare, but appear to be present whenever clusters form in large numbers. While bound star clusters may generally form with a higher efficiency in environments of high star formation rate, many of the apparent differences between clusters in starbursts and ``normal'' galaxies might be simply due to sampling effects. ", "introduction": "It may be worth recalling some of the main properties of the Milky Way open cluster system. The census of open clusters is still highly incomplete beyond distances of a few kpc from the Sun, although the situation is improving with new surveys such as 2MASS (see e.g.\\ the contributions by Carpenter and Hanson in this volume). The luminosity function of Milky Way open clusters was analysed by \\citet{van84}, who found it to be well modelled by a power-law $N(L)dL \\propto L^{-1.5}dL$ over the range $-8 < M_V < -3$. However, they also noted that extrapolation of this luminosity function would predict about 100 clusters as bright as $M_V=-11$ in the Galaxy, clearly at odds with observations, and thus suggested some flattening of the LF slope at higher luminosities. The brightest known young clusters (e.g.\\ NGC~3603, $h$ and $\\chi$ Per) have absolute $V$ magnitudes of $M_V\\sim-10$, corresponding to total masses of several thousand \\msun . Recently, there have been claims that the Cyg OB2 association might be an even more massive cluster \\citep{kno00}, but this object is probably too diffuse to be a bound star cluster (though it does have a compact core). There are, however, a number of old ($>1$ Gyr) open clusters in the Milky Way with masses of $\\sim10^4$ \\msun\\ \\citep{friel95}. These objects are likely to have lost a significant fraction of their total mass over their lifetimes, and may thus originally have been even more massive. They serve to illustrate that, even in the Milky Way, the distinction between globular and open clusters is not always clear-cut. It has been recognized for about a century that the Magellanic Clouds, and the LMC in particular, host a number of ``blue globular clusters'' \\citep{shap30}. % Among the most massive of these is NGC~1866, with a mass of around $10^5 \\, \\msun$ and an age of $\\approx100$ Myr \\citep{fis92,van99}. An older example is NGC~1978 with similar mass but an age of 2--3 Gyr, clearly demonstrating that at least some such clusters can survive for several Gyrs. The interaction between the LMC and the Milky Way has probably affected the star formation history of the LMC, which is known to be bursty with major peaks in the star formation rate correlating with perigalactic passages \\citep{smeck02}. One might argue, then, that the formation of YMCs in the LMC could be induced by interaction with the Milky Way. However, the LMC is not the only example even in the Local Group of a galaxy that hosts YMCs. Another well-known example is M33, which does not display evidence for a bursty cluster (and, presumably star-) formation history \\citep{cs82,cs88}. \\citet{chan99,chan01} have identified many more star clusters in this galaxy, though not all are particularly massive. With the launch of HST it became possible to investigate more crowded and/or distant systems in detail and attention started to shift towards more extreme starbursts, including a large number of merger galaxies (e.g.\\ Whitmore, this volume). It is now clear that luminous, young star clusters often form in very large numbers in such galaxies, and this has led to suggestions that formation of ``massive'' star clusters might require special conditions such as large-scale cloud-cloud collisitions \\citep{js92}. However, the question remains to be answered why some non-interacting galaxies also contain YMCs, whereas apparently the Milky Way does not. YMCs are now being found in an increasing number of non-interacting galaxies, posing a severe challenge for formation scenarios which require special conditions. ", "conclusions": "It is becoming increasingly clear that ``massive'' star clusters can form in a wide variety of galaxies, and not just in mergers or otherwise disturbed galaxies. With the possible exception of some dwarf galaxies, the luminosity distributions of young star clusters generally appear to be power-laws. If cluster luminosities are sampled at random from a power-law distribution, the most luminous clusters will naturally be rare, but so far there is no evidence for a statistically significant upper cut-off. In other words, very luminous (and massive) clusters appear to form whenever clusters form in large numbers. % This is illustrated by the fact that young star clusters with masses up to $\\sim10^6 \\, \\msun$ have been identified in the disks of several apparently normal, isolated spiral galaxies with rich cluster systems. These galaxies, such as NGC~5236, NGC~6946 are characterised by high star formation rates, but these do not generally appear to be triggered by interactions with other galaxies. Dynamical mass estimates are now available for a small number of these clusters, and the mass-to-light ratios are compatible with standard Kroupa-type IMFs. There is every reason to be optimistic that important clues to the formation of classical globular clusters may be obtained by studying their younger counterparts in the Local Universe." }, "0403/astro-ph0403134_arXiv.txt": { "abstract": "{We discuss the lifetimes and evolution of dense cores formed as turbulent density fluctuations in magnetized, isothermal molecular clouds. We consider numerical simulations in which we measure the cores' magnetic criticality and Jeans stability in relation to the magnetic criticality of their ``parent clouds'' (the numerical boxes). In subcritical boxes, dense cores do not form, and collapse does not occur. In supercritical boxes, some cores collapse, being part of larger clumps that are supercritical from the start, and whose central, densest regions (the cores) are initially subcritical, but rapidly become supercritical, presumably by accretion along field lines. Numerical artifacts are ruled out. The time scales for cores to go from subcritical to supercritical and then collapse are a few times the core free-fall time, $\\tfc$. Our results suggest that cores are out-of-equilibrium, transient structures, rather than quasi-magnetostatic configurations. } \\resumen{Discutimos los tiempos de vida y la evoluci\\'on de los n\\'ucleos densos formados como fluctuaciones turbulentas de densidad en nubes moleculares isot\\'ermicas magnetizadas. Consideramos simulaciones num\\'ericas en las que medimos la criticalidad magn\\'etica y la estabilidad de Jeans de los n\\'ucleos, en relaci\\'on a la criticalidad magn\\'etica de sus nubes ``madre'' (los dominios de integraci\\'on num\\'erica). En dominios subcr\\'iticos, no se forman n\\'ucleos densos, y no ocurre colapso gravitacional. En dominios supercr\\'iticos, algunos n\\'ucleos se colapsan, formando parte de grumos m\\'as grandes que son supercr\\'iticos desde el inicio, y cuyas regiones m\\'as densas (los n\\'ucleos) son inicialmente subcr\\'iticas, pero r\\'apidamente se tornan supercr\\'iticos, presumiblemente por acreci\\'on a lo largo de las l\\'ineas de campo magn\\'etico. Descartamos la posibilidad de artefactos num\\'ericos. Las escalas de tiempo en las que los n\\'ucleos transitan desde el estado subcr\\'itico, pasando por el supercr\\'itico, hasta finalmente el colapso, son de unas cuantas veces su tiempo de ca\\'ida libre, $\\tfc$. Nuestros resultados sugieren que los n\\'ucleos son estructuras transientes, fuera de equilibrio, y no configuraciones cuasi-magneto-hidrost\\'aticas. } \\addkeyword{ISM: Molecular clouds} \\addkeyword{ISM: Turbulence} \\addkeyword{Stars: Formation} \\begin{document} ", "introduction": "\\label{sec:intro} The prevailing view (which we hereafter refer to as the ``standard (magnetic support) model'' of star formation; see, e.g., the reviews by Shu, Adams \\& Lizano 1987; McKee et al.\\ 1993) concerning low-mass-star-forming clumps is that they are quasi-static equilibrium configurations with so-called ``subcritical'' mass-to-magnetic-flux ratios, so that the clumps are supported against their self-gravity by the magnetic field in the direction perpendicular to it, and by a combination of thermal and turbulent pressures along the field. Under ideal MHD conditions, the magnetic field is ``frozen'' into the plasma, and the magnetic flux is conserved. Under the additional assumption that the clump's mass is also constant, then the mass-to-flux ratio is a fixed parameter of the clump, which therefore cannot collapse if this ratio is subcritical (meaning that the core's self-gravity is never enough to overwhelm the magnetic support). However, because the cold molecular gas is only partially ionized, the process known as ambipolar diffusion causes a loss of magnetic flux from the clumps on time scales long compared to their free-fall time, allowing them to contract and form denser cores that will ultimately collapse. However, it well known that molecular clouds are supersonically turbulent (e.g., Larson 1981; Blitz \\& Williams 1999), and it is becoming increasingly accepted that the cores within them are the density fluctuations induced by the turbulence (\\BP, \\VS\\ \\& Scalo 1999; see also the reviews by \\VS\\ et al.\\ 2000; Mac Low \\& Klessen 2004). In this context, the cores have a highly dynamical origin (supersonic compressions), and their masses are hardly fixed. Moreover, it is natural to ask whether they can settle into hydrostatic equilibria, an event which requires the equilibria to be stable (or ``attracting'', in the language of nonlinear phenomena). Otherwise, the dynamic density fluctuations will just ``fly past'' the equilibrium state on their way to collapse, or else ``rebound'' and merge back with their environment, if they do not quite reach the equilibrium point. In this case, the cores' lifetimes should be much shorter than in the standard model, probably comparable to their free-fall times. In the present contribution, we argue in favor of this scenario. To this end, we discuss the formation and evolutionary time scales of cores that form in numerical simulations of isothermal, compressible MHD turbulence, in relation to the magnetic criticality of the whole computational box. Here we present a brief overview. For a full, detailed discussion, see \\VS\\ et al.\\ (2004). ", "conclusions": "\\label{sec:conclusions} In this contribution, we have presented a discussion and numerical simulations of the formation, nature, and lifetimes of dense cores in magnetized clouds. We argued that the mass-to-flux ratio of a cloud puts an upper bound to that of any clump or core within it, and so the only way to form supercritical cores is within supercritical clouds (neglecting ambipolar diffusion). Moreover, numerical simulations of marginally subcritical clouds show that no gravitationally bound cores form in this case, while in supercritical clouds the gravitationally bound cores that form occur inside clumps that are supercritical, and rapidly become supercritical themselves and collapse, in timescales that do not exceed twice their local free-fall time $\\tfc$. We also showed that the mass-to-flux ratio of structures defined through a density threshold (as would be the case of cores observed in a single molecular line, which requires the density to be larger than a certain value in order to be excited) does not remain constant, because the core is continuously connected to its parent structure, from which it can accrete mass. Our results support the notion that clumps and cores are out-of-equilibrium, transient structures, and that a class of ``failed'' cores should exist, that will not form stars." }, "0403/physics0403086_arXiv.txt": { "abstract": "The issue of asymmetric uncertainties resulting from fits, nonlinear propagation and systematic effects is reviewed. It is shown that, in all cases, whenever a published result is given with asymmetric uncertainties, the {\\it value} of the physical quantity of interest {\\it is biased} with respect to what would be obtained using at best all experimental and theoretical information that contribute to evaluate the combined uncertainty. The probabilistic solution to the problem is provided both in exact and in approximated forms. ", "introduction": "We often see published results in the form $$\\mbox{`best value'}\\ ^{+\\Delta_+}_{-\\Delta_-}\\,,$$ where $\\Delta_+$ and $\\Delta_-$ are {\\it usually} positive.\\footnote{For examples of measurements having $\\Delta_+$ and $\\Delta_-$ with all combinations of signs, see public online tables of Deep Inelastic Scattering results.\\cite{HERA} I want to make clear since the very beginning that it is not my intention to blame experimental or theoretical teams which have reported in the past asymmetric uncertainty, because we are all victims of a bad tradition in data analysis. At least, when asymmetric uncertainties have been given, there is some chance to correct the result, as described in Sec.~\\ref{sec:thumb}. Since some asymmetric contributions to the global uncertainties almost unavoidably happen in complex experiments, I am more worried of collaborations that never arrive to final asymmetric uncertainties, because I must imagine they have symmetrised somehow the result but, I am afraid, without applying the proper shifts to the `best value' to take into account asymmetric contributions, as it will be discussed in the present paper.} As firstly pointed out in Ref.~\\cite{ConMirko} and discussed in a simpler but more comprehensive way in Ref.~\\cite{BR}, this practice is far from being acceptable and, indeed, could bias the believed value of important physics quantities. The purpose of the present paper is, summarizing and somewhat completing the work done in the above references, to remind where asymmetric uncertainty stem from and to show why, as they are usually treated, they bias the value of physical quantities, either in the published result itself or in subsequent analyses. Once the problems are spotted, the remedy is straightforward, at least within the Bayesian framework (see e.g. \\cite{BR}, or \\cite{RPP_GdA} and \\cite{RPP_Dose} for recent reviews). In fact the Bayesian approach is conceptually based on the intuitive idea of probability, and formally grounded on the basic rules of probability (what are usually known as the probability `axioms' and the `conditional probability definition') plus logic. Within this framework many methods of `conventional' statistics are reobtained, as approximations of general solutions, under well stated conditions of validity. Instead, in the conventional, frequentistic approach {\\it ad hoc} formulae, prescriptions and un-needed principles are used, often without understanding what is behind these methods -- before a `principle' there is nothing! The proposed Bayesian solutions to cure the troubles produced by the usual treatment of asymmetric uncertainties is to step up from approximated methods to the more general ones (see e.g. Ref.~\\cite{BR}, in particular the top down approximation diagram of Fig.~2.2). In this paper we shall see, for example, how $\\chi^2$ and minus log-likelihood fit `rules' can be derived from the Bayesian inference formulae as approximated methods and what to do when the underlying conditions do not hold. We shall encounter a similar situation regarding standard formulae to propagate uncertainty. Some of the issues addressed here and in Refs. \\cite{ConMirko} and \\cite{BR} have been recently brought to our attention by Roger Barlow~\\cite{Barlow}, who proposes frequentistic ways out. Michael Schmelling had also addressed questions related to `asymmetric errors', particularly related to the issue of weighted averages~\\cite{Schmelling}. The reader is encouraged to read also these references to form his/her idea about the spotted problems and the proposed solutions. In Sec. \\ref{sec:propagation} the issue of propagation of uncertainty is briefly reviewed at an elementary level (just focusing on the sum of uncertain independent variables -- i.e. no correlations considered) though taking into account asymmetry in probability density functions (p.d.f.) of the {\\it input} quantities. In this way we understand what `might have been done' (we are rarely in the positions to exactly know ``what has been done'') by the authors who publish asymmetric results and what is the danger of improper use of such a published `best value' -- {\\it as is} -- in subsequent analyses. Then, Sec.~\\ref{sec:sources} we shall see in where asymmetric uncertainties stem from and what to do in order to overcome their potential troubles. This will be done in an exact way and, whenever is possible, in an approximated way. Some rules of thumb to roughly recover sensible probabilistic quantities (expected value and standard deviation) from results published with asymmetric uncertainties will be given in Sec.~\\ref{sec:thumb}. Finally, some conclusions will be drawn. ", "conclusions": "Asymmetric uncertainties do exist and there is {\\it no way to remove them artificially}. If they are not properly treated, i.e. using prescriptions that do not have a theoretical ground but are more or less rooted in the physics community, the published result is biased. Instead, if they are properly treated using probability theory, in most cases of interest the final result is practically symmetric and approximately Gaussian, with expected value and standard deviations which take into account the several shifts due to individual asymmetric contributions. Note that some of the simplified methods to make statistical analyses had a {\\it raison d'\\^etre} many years ago, when the computation was a serious limitation. Now it is not any longer a problem to evaluate, analytically or numerically, integrals of the kind of those appearing e.g. in Eqs.(\\ref{eq:prop_general}), (\\ref{eq:theta_E}) and (\\ref{eq:theta_sigma}). In the case the final uncertainty remains asymmetric, the authors should provide detailed information about the `shape of the uncertainty', giving also most probable value, probability intervals, and so on. But the best estimate of the {\\it expected value and standard deviation should be always given} (see also the {\\it ISO Guide}~\\cite{ISO}). To conclude, I would like to leave the final word to my preferred quotation with whom I like to end seminars and courses on probability theory applied to the evaluation and the expression of uncertainty in measurements: \\begin{quote} {\\sl \\small ``Although this {\\rm Guide} provides a framework for assessing uncertainty, it cannot substitute for critical thinking, intellectual honesty, and professional skill. The evaluation of uncertainty is neither a routine task nor a purely mathematical one; it depends on detailed knowledge of the nature of the measurand and of the measurement. The quality and utility of the uncertainty quoted for the result of a measurement therefore ultimately depend on the understanding, critical analysis, and integrity of those who contribute to the assignment of its value.''}\\cite{ISO} \\end{quote} \\vspace{1.0cm} \\noindent It is a pleasure to thank {\\it Superfaber} (Fabrizio Fabbri in {\\it hepnames}) for helpful discussions on the subject and for his {\\it super}vision of the manuscript." }, "0403/astro-ph0403302_arXiv.txt": { "abstract": "{ We present high resolution optical spectroscopy of three candidate members of the Taurus-Auriga star forming region. Based on the spectral type, the strength, profile and width of the H$\\alpha$ line, the detected lithium at 6708 \\AA, the location of these objects in a H-R diagram and the comparison with similar objects belonging to young stellar associations, we determine that they are bona fide members of the SFR, with about $\\sim$3 Myr, have masses at or below the substellar limit and, at least in one case, there is active accretion from a circum(sub)stellar disk. This result suggests that high mass brown dwarfs go through a Classical TTauri phase and form like stars, from colapse and fragmentation of a molecular cloud. ", "introduction": "Brown dwarfs, objects unable to fuse hydrogen in an stable manner (i.e., with masses below at about 0.072 M$_\\odot$, Baraffe et al. 1998), pose an important problem to the theory of stellar formation. Several formation mechanisms have been proposed, including formation like a star (from collapse and fragmentation of a molecular cloud, Padoan P., \\& Nordlund 2004) to a planet-like process (from a circumstellar disk) or as stellar ``embryos'', ejected from multiple systems before they are able to accrete enough matter (Reipurth \\& Clarke 2001; Bate et al. 2002). These proposed mechanisms have different implication in their formation, their evolution and their properties. In particular, if brown dwarfs are created like stars, it seems that they should go through an phase of active accretion from a circum(sub)stellar disk, such as low mass stars do (Shu et al$.$ 1987). Actually, since the last couple of years, different groups have presented indirect and direct evidences of active accretion in a handful of young brown dwarfs belonging to several nearby star forming regions (SFR), open clusters and moving groups (Fern\\'andez \\& Comer\\'on 2001; Muench et al. 2001; Natta \\& Testi 2001; Natta et al. 2001, 2002; Testi et al. 2002; Jayawardhana et al. 2002ab, 2003ab; Muzerolle et al. 2003; Barrado y Navascu\\'es et al. 2002, 2003, 2004; Barrado y Navascu\\'es \\& Mart\\'{\\i}n 2003; Mohanty \\& Basri 2003; Mohanty, Jayawardhana \\& Barrado y Navascu\\'es 2003; Comer\\'on et al. 2003) In this paper, we enlarge the sample by collecting high resolution spectroscopy of three proposed members of the Taurus-Auriga SFR, located at 140 pc and about 1-3 Myr (for an update on Taurus, see Luhman et al. 2003 and references therein). ", "conclusions": "\\subsection{Spectral types} Spectral types were derived by comparing with several spectral templates, by using the order around 7050 \\AA, corresponding to a TiO band, following Mohanty et al. (2004), since this range is very sensitive to effective temperature. Errors can be estimated as half a subclass. Our values are very close to those obtained by Brice\\~no et al. (2002) from low resolution spectra (see Table 1). \\setcounter{figure}{0} \\begin{figure} \\centering \\includegraphics[width=8.2cm]{barradoF1.ps} \\caption{ Spectra around HeI6678 \\AA{} and LiI6707.8 \\AA. } \\end{figure} \\subsection{On the lithium abundance, age, and mass} We have detected lithium 6707.8 \\AA{} in two out of the three targets (KPNO-Tau-05 and KPNO-Tau-08, Figure 1). The spectrum of KPNO-Tau-03 has worse quality. With some caveats, the visual inspection indicates that this feature is also present. Note that Brice\\~no et al. (2002) states that they detected lithium with their low resolution spectrum (at higher signal-to-noise ratio). This element is easily destroyed in the stellar interior, being its surface abundance dependent on mass and age (as well as other second order parameters). In fact, brown dwarfs which are more massive than about 0.060 M$_\\odot$ do deplete lithium in a time scale of few tens of million years (D'Antona \\& Mazzitelli 1994; Burrows et al. 1997; Baraffe et al. 1998). The Taurus-Auriga complex has an age between 1 and 3 Myr. Since our three candidate members have spectral types between M6 and M7, they should have masses equal or larger than 0.06 M$_\\odot$, according to Baraffe et al. (1998). Therefore, the detection of this alkali clearly indicates that these objects are in the pre-Main Sequence (PMS). Moreover, from the statistical point of view, the likelihood of having three late-M, PMS interloper whose spectral and photometric properties coincide with the Taurus sequence is negligible. Therefore, we have to conclude that they, indeed, belong to the association. \\setcounter{figure}{1} \\begin{figure} \\centering \\includegraphics[width=8.2cm]{barradoF2.ps} \\caption{ Lithium equivalent width versus the spectral type. The solid line corresponds to the upper envelope of the values measured in young open clusters. The long-dashed line delimits the areas for weak-line and post-T~Tauri stars (adapted from Mart\\'{\\i}n 1997 and Mart\\'{\\i}n \\& Magazz\\`u 1999). Short-dashed and dotted lines correspond to the cosmic abundances --A(Li)=3.1-- from gravities of Logg=4.5 and 4.0, respectively (curves of growth from Zapatero Osorio et al$.$ 2002). All Pleiades and IC2391 members with measured lithium equivalent width are shown as open circles and triangles --upper limits--. Note the lithium depletion boundary at $\\sim$M5.5. Sigma Orionis low mass stars and brown dwarfs appear as solid squares (Zapatero Osorio et al$.$ 2002). The big asterisks represent KPNO-Tau-3, 5 and 8. } \\end{figure} Figure 2 displays values of the lithium equivalent width (W) measured for our targets (large asterisks), pre-main sequence members of the $\\sim$5 Myr Sigma Orionis cluster (Zapatero Osorio et al. 2002, solid squares), and stellar and substellar members of IC2391 --53 Myr-- and the Pleiades --125 Myr-- from Barrado y Navascu\\'es et al. (1999, 2004), Soderblom et al. (1993), Garc\\'{\\i}a-L\\'opez et al. (1994), Jones et al. 1996, Stauffer et al. (1998) and Jeffries et al. (1999). The solid line describe the maximum W(Li) measured in cluster stars. The long-dashed curve is an update version of the criterion used by Mart\\'{\\i}n (1997) and Mart\\'{\\i}n \\& Magazz\\`u (1999) to distinguish between Weak-line and post- TTauri stars. Finally, the dotted and short dashed lines correspond to the Zapatero Osorio et al. (2002) curves of growth for Log\\,g=4.0 and 4.5 and an abundance of A(Li)=3.1 (i.e., cosmic abundance). We have labeled in the plot different sections. The W(Li) for KPNO-Tau-5 is lower than the measured values in the other two Taurus objects. This fact might be due to the presence of optical veiling, about r(6700)$\\sim$0.2, although we do not think this is the case. On one hand, no other sign of accretion has been found in this object (see next section). Moreover, this equivalent width is compatible with the dispersion found in Sigma Orionis and the --undepleted- values characteristic of cluster brown dwarfs. For the other two objects, no veiling seems to be present either. In the lithium equivalent width dispersion is real both in Taurus and Sigma Orionis cluster (the same effect is present in the Lambda Orionis association, Barrado y Navascu\\'es, Stauffer \\& Jayawardhana 2004), it might imply diferences in the surface abundances due to additional mixing mechanisms in the stellar and substellar interior, different to pure convection (such as those proposed in solar mass stars), differences in the structure due to different stellar and substellar parameters (such as fast rotation in the case of pre-Main sequence stars, Mart\\'{\\i}n \\& Claret 1996), or to induced effects related to activity and/or rotation over the spectral feature itself or the surrounding continuum (see the discusion in Barrado y Navascu\\'es et al. 2001 and references therein). As a summary, based on these data, mainly from the detection of lithium and accretion when present (see next section), we can conclude that these three objects are young and, indeed, belong to the Taurus region. In the case of KPNO-Tau-5, due to its spectral type (M7) its substellar nature is well established. The other two are located at the substellar borderline and their nature is not so firmly established due to uncertainties in the models and the spectral type determination. Once membership to the stellar association has been proved, we have derived the bolometric magnitudes from $Ic$ and $Ks$, the distance modulus $(m-M)_0$=5.731 (140 pc, Kenyon, Dobrzycka \\& Hartmann 1994), the reddening derived by Brice\\~no et al. (2002) and the bolometric corrections by Comer\\'on et al. (2000) and Tinney et al. (1993) for these two bands ($Ic$ and $Ks$, respectively). Masses were computed based on models by Baraffe et al. (1998). Effective temperatures were obtained using several scales, namely Luhman (1999) for intermediate gravity and Leggett (2000, 2001). All the measured and derived values are listed in Table 1. \\setcounter{figure}{2} \\begin{figure} \\centering \\includegraphics[width=8.2cm]{barradoF3.ps} \\caption{HR diagram of KPNO-Tau-3 and 8 (on top of each other) and KPNO-Tau-5. The values represented by asterisks and open stars were derived from $Ic$ and $Ks$ magnitudes, respectively. } \\end{figure} Figure 3 displays a HR diagram. The isochrones and evolutionary tracks --solid and dashed lines, respectively-- are from Baraffe et al. (1998). Asterisks and open stars correspond to bolometric luminosities obtained from $Ic$ and $Ks$, respectively. For this particular diagram, we made use of the effective temperature scale by Luhman (1999). Other temperature scales, such as that from Leggett et al. (2001), would shift the location of these three objects (two of them are almost on top of each other, the small shift in Teff is arbitrary) to the right hand-side, making them younger and less massive. Note, however, that Luhman' scale has been tunned specifically for the model we are using here. In any case, regardless the election of models (alternative models are, for instance, Burrows et al. 1997; D'Antona \\& Mazzitelli 1994, 1997, 1998; Chabrier et al. 2000; Baraffe et al. 2002), bolometric corrections and effective temperature scale, these members are at or below the substellar frontier and have an age of about 3 Myr. \\setcounter{figure}{3} \\begin{figure} \\centering \\includegraphics[width=8.2cm]{barradoF4.ps} \\caption{Color-color diagrams. The position of the Taurus BD candidates are indicated with large asterisks. Crosses indicate the position of classical TTauri stars belonging to Orion stellar population (Herbig \\& Bell 1988). The thick-solid and dashed lines correspond to the locii of the main sequence stars (from Bessell \\& Brett 1988; Kirkpatrick et al$.$ 2000; Leggett et al. 2001) and CTT stars (Meyer et al$.$ 1997, Barrado y Navascu\\'es et al. 2003), respectively. } \\end{figure} \\subsection{Near infrared photometry, H$\\alpha$ emission and accretion} None of our three targets seems to have neither the presence of forbidden lines, characteristics of outflows, nor near infrared infrared excesses, coming from an accretion disk. Figure 4 compares the colors $(Ic-J)$ and $(H-Ks)$, and includes the loci for dwarfs --Bessell \\& Brett 1988; Kirkpatrick et al. 2000; Leggett et al. 2001)-- and Classical TTauri stars (Meyer et al. 1997, Barrado y Navascu\\'es et al. 2003) and Classical TTauri stars from Orion (Herbig \\& Bell 1988). As can be seen, the Taurus objects have photometric properties --data from 2MASS-- similar to those MS stars of similar spectral type and, as stated in the previous paragraph, no near IR excess is seen. This fact, by itself, cannot prove or disprove the presence of a circum(sub)stellar disk, since a hole can be present or the disk temperature can be very cold. Moreover, the IR excess might depend on the orientation of the disk. See, for example, the case of LS-RCrA~1, M6.5 brown dwarf belonging to RCrA dark cloud (Barrado y Navascu\\'es, Mohanty \\& Jayawardhana 2004). Additional observations and longer wavelengths --more sensitive to cooler disks- both from ground-based telescopes and space-borne instruments such as those in Spitzer Space Telescope, can help to shed some light in this issue. \\setcounter{figure}{4} \\begin{figure} \\centering \\includegraphics[width=8.2cm]{barradoF5.ps} \\caption{ H$\\alpha$ profiles. The instrumental profile is included as a dashed line (top spectrum). } \\end{figure} We have detected and measured H$\\alpha$ in emission for these three objects. The profiles of this feature are displayed in Figure 5. Note the possible asymmetry in KPNO-Tau-8, the double peak in KPNO-Tau-5 and KPNO-Tau-3, and the width of the line (310 km/s), typical of accreting brown dwarfs (White \\& Basri 2003; Jayawardhana et al. 2003). In fact, this last object is above the criterion defined by Barrado y Navascu\\'es \\& Mart\\'{\\i}n (2003) which discriminate between accreting and non accreting objects, as Figure 6 clearly indicates. We note, however, that this criterion depends on low resolution spectroscopy, which normally yields larger equivalent widths compared with higher resolution data. In this diagram, we have included Classical and Weak-line TTauri stars belonging to Taurus as solid and open circles. Members without classification are included as crosses. Our three targets appear as large asterisks. Overlapping squares and big circles denote those members with forbidden emission lines and near-infrared excesses, respectively. This large W(H$\\alpha$) agree with the fact that we also detect HeI6678 \\AA, another accretor indicator (see Figure 1). \\setcounter{figure}{5} \\begin{figure} \\centering \\includegraphics[width=8.2cm]{barradoF6.ps} \\caption{H$\\alpha$ equivalent widths for members of the Taurus SFR. Solid and open circles correspond to Classical and TTauri stars. Objects with no classification are shown as crosses. Large open circles or square represent objects with mid-IR excesses and forbidden lines, respectively. The three objects studied here are displayed as large asterisks. The dotted, bold curve is the saturation criterion, whereas two previously proposed criteria (5 and 20 \\AA) to separate CTTS and WTTS are included as long-dashed, thin horizontal segments. The vertical dotted segment denotes the location of the substellar frontier. } \\end{figure} Active accretion in at least one Taurus member whose mass is close or below the substellar limit is important for several reasons. First, although the sample studied in this paper is very small, it suggests that a significant fraction of the very low mass stars and high mass brown dwarfs might harbor an accretion disk. Second, the new data amass additional evidence for accretion in the substellar domain. To the best of our knowledge, there are only three papers dealing with high resolution spectra in Taurus brown dwarfs (White \\& Basri 2003; Muzerolle et al. 2003; and Jayawardhana et al. 2003). The first work found three accretors in a sample of ten very low mass stars and brown dwarfs, although the less massive, a M6.5 (GM~Tau), has a mass just above the substellar limit. The second paper includes four objects whose spectral type is M6 or M7. None of them seem to undergo active accretion based on the width of H$\\alpha$, although the authors claim that one of them, namely MHO-5, is accreting based on the detection of forbidden lines of oxygen at 6300 and 6363 \\AA{ } and CaII IRT. Moreover, three out of the four were observed at moderate resolution, including MHO-5 (R$\\sim$8000). Regarding the later study, the four M7-M7.75 brown dwarfs discussed there, with masses down to 0.05 M$_\\odot$ (again, using Baraffe et al. 1998 models), do not show accretion either. Muzerolle et al. (2003) also analyzed seven Taurus members whose spectral types are M4.75--M5.75 (just above or at the substellar borderline). Three out of these seven are accreting, based on the width of H$\\alpha$ and other spectral indicators. Therefore, up to date, our study presents the only brown dwarf belonging to Taurus (kpno-tau-03), at the substellar limit, which has been proved to be accreting. In conjunction with the studies quoted in the previous paragraph, our results indicate that about 10\\% of the Taurus brown dwarfs (one out of 11) is actively accreting. This fraction is much smaller than the accretion occurrence in low mass stars in the association or the estimate for the substellar domain (about 50\\%) based purely on the strength of H$\\alpha$ measured in low resolution spectra (Barrado y Navascu\\'es \\& Mart\\'{\\i}n 2003). The statistical criterion defined in this last work is based on the saturation of the activity, as measured in several young open clusters (namely IC2391, Alpha Per and the Pleiades, with ages ranging from about 50 to 125 Myr). The discrepancy in the fraction of accreting brown dwarfs might imply that there is an additional source of flux in H$\\alpha$ line or that the accreting criterion based on the H$\\alpha$ width at 10\\% of the maximum intensity, as defined by Jayawardhana et al. (2003), i.e. 200 km/s, is too restrictive. The detection of accretion in substellar objects indicates that they undergo a phase similar to Classical TTauri stars (for a review, see Bertout 1989 or Appenzeller \\& Mundt 1989). A handful of other brown dwarfs belonging to other young stellar association, with ages ranging from 1 to 10 Myr, have been observed with high resolution spectroscopy and the presence of accretion confirmed. In at least one, LS-RCrA~1 (Fern\\'andez \\& Comer\\'on 2001; Barrado y Navascu\\'es, Mohanty \\& Jayawardhana 2004), outflows have been seen, by means of the detection of intense, narrow forbidden lines. In a previous paper we have named them Classical TTauri substellar analogs (CTTSA). Therefore, we can conclude, at least in the case of high mass brown dwarfs (Mas$\\sim$0.072--0.04 M$_\\odot$), that they are formed as low mass stars, by fragmentation and collapse of the original molecular cloud. Of course, additional studies are needed to confirm this preliminary conclusion, in particular high resolution imaging in the near and mid- infrared, as well as in the optical. Narrow band imaging might show whether these objects present outflows similar to those observed in some Classical TTauri stars." }, "0403/astro-ph0403628_arXiv.txt": { "abstract": "{This paper presents the VIMOS VLT Deep Survey around the Chandra Deep Field South (CDFS). We have measured 1599 new redshifts with VIMOS on the European Observatory Very Large Telescope - UT3, in an area $21 \\times 21.6$ arcmin$^2$, including 784 redshifts in the Hubble Space Telescope - Advanced Camera for Surveys GOODS area. 30\\% of all objects with $I_{AB}=24$ have been observed independently of magnitude, indicating that the sample is purely magnitude limited. We have reached an unprecedented completeness level of 88\\% in terms of the ratio of secure measurements vs. observed objects, while 95\\% of all objects have a redshift measurement. A total of 1452 galaxies, 139 stars, 8 QSOs have a redshift identification, 141 of these being unsecure measurements. The redshift distribution down to $I_{AB}=24$ is peaked at a median redshift z=0.73, with a significant high redshift tail extending up to $\\sim4$. Several high density peaks in the distribution of galaxies are identified. In particular, the strong peak at z=0.735 contains more than 130 galaxies in a velocity range $\\pm2000$km/s distributed all across the transverse $\\sim$20 $h^{-1}$ Mpc of the survey. We are releasing all redshifts to the community, along with the cross identification with HST-ACS GOODS sources on the CENCOS database environment http://cencosw.oamp.fr. ", "introduction": "Understanding the major steps in the evolution of galaxies still remains a major challenge to modern astrophysics. While the general theoretical framework of the hierarchical growth of structures in the universe including the build up of galaxies is well in place (e.g. \\cite{peacock04}), at high redshifts this picture remains largely unconstrained by observations. The detailed properties of the main population of galaxies from large samples representative of the universe at various epochs remain to be established across most of the life of the universe beyond the large local volumes explored by the 2dFGRS (\\cite{colless} and the SDSS (\\cite{SDSS}), and expanding from smaller exploratory surveys (\\cite{lilly95}, \\cite{lefevre95}, \\cite{steidel03}, \\cite{cimatti03}). The VIMOS VLT Deep Survey (VVDS) is a deep redshift survey aimed at studying the evolution of galaxies, large scale structures and AGNs over more than 90\\% of the current age of the universe. The unique feature of the VVDS is the simple magnitude selection applied to define a complete magnitude limited sample of distant galaxies, with a goal of more than 100000 objects observed in multi-object spectroscopy. The VVDS rests on the observations of 5 different fields to smooth out the effects of cosmic variance when building the statistical properties of the galaxy population (\\cite{lefevre04}). This paper presents the redshift survey observations of 1599 objects with $I_{AB}\\leq24$ conducted by the VVDS team around the Chandra Deep Field South, including the HST-GOODS area (\\cite{goods}). The observations have been carried out with the VIsible Multi-Object Spectrograph (VIMOS) on the 8.2m Melipal telescope of the European Southern Observatory Very Large Telescope. We are describing the processing steps and redshift measurements, and the associated quality control we have applied to these data. The content of the final catalog is detailed, as well as the cross identification with the Hubble Space Telescope Advanced Camera for Surveys GOODS images, and we present the main entries available from our interactive database. The main properties of the sample are briefly presented, including the redshift distribution of the sample. ", "conclusions": "In the framework of the VIMOS VLT Deep Survey (VVDS), we have observed a large sample of galaxies around the Chandra Deep Field South, and are releasing the redshift data to the community. A total of 1599 objects with $I_{AB} \\leq 24$ have a measured redshift. The completeness in redshift measurement for the targeted objects is high, above 88\\%. We find that the redshift distribution has a median of $z=0.73$, with strong high density peaks observed across the field. The combination of this redshift survey and the HST-ACS GOODS survey enables detailed studies of the evolution of galaxies in the Chandra Deep Field South." }, "0403/astro-ph0403673_arXiv.txt": { "abstract": "Abell~399 and Abell~401 are both rich clusters of galaxies, with global temperatures of 7.2~keV (Abell~399) and 8.5~keV (Abell~401) respectively. They lie at a projected separation of $\\sim$3~Mpc, forming a close pair. We have observed the system with the \\xmm satellite. The data of each cluster show significant departures from our idealised picture of relaxed rich clusters. Neither of the two contains a cooling flow, and we find that their central regions are nearly isothermal, with some small-scale inhomogeneities. The image analysis derives $\\beta$-values that are smaller than the canonical value of 0.65, and the surface brightness distribution is not symmetric around the central cD galaxies: there are irregularities in the central $\\sim$200~kpc, and asymmetries on larger scales, in that the intracluster gas in each cluster is more extended towards the other member of the system. Both clusters host extended radio halos and a plethora of tailed radio galaxies. The halo in Abell~399 appears to be correlated with a sharp edge apparent in the \\xmm images, and a region of harder X-ray emission. There is also evidence for enhanced X-ray flux in the region between the two clusters, where the temperature is higher than our expectations. Although tidal or compression effects might affect the large scale structure of the two clusters, we show that these cannot account for the distortions seen in the inner regions. We argue that the reasonably relaxed morphology of the clusters, and the absence of major temperature anomalies, argues against models in which the two have already experienced a close encounter. The properties of the intermediate region suggests that they are at an early stage of merging, and are currently interacting mildly, because their separation is still too large for more dramatic effects. The substructure we find in their inner regions seems to point to their individual merging histories. It seems likely that in the Abell~399/401 system, we are witnessing two merger remnants, just before they merge together to form a single rich cluster of galaxies. This picture is consistent with recent numerical simulations of cluster formation. ", "introduction": "It is now common wisdom that clusters of galaxies are formed hierarchically, by the merging of smaller mass units, preferentially along large-scale filaments. During such violent events one expects physical processes to take place that have significant impact on the properties of the constituents of clusters (i.e., gas, galaxies), that would define their subsequent evolution, as well as the energy and entropy budget in the largest structures in the Universe. During mergers, shock waves propagate within the intracluster medium (ICM), thus increasing its thermal energy and entropy. The cluster galaxies may be severely stripped, leaving their metal rich interstellar media (ISM) behind in the cluster, contributing in this way to the enrichment of the ICM (e.g., Acreman et al. 2003). Cooling flows may get displaced or destroyed entirely (e.g., G\\'{o}mez et al. 2002) . \\begin{figure} \\begin{center} \\leavevmode \\epsfxsize 1.05\\hsize \\epsffile{m1m2_all_500-10000.ps} \\caption{A mosaic image of all 4 pointings in the (0.5-10)~keV energy range. Background subtracted images from the two MOS instruments (MOS1 and MOS2) are superimposed. The pixel size is 8~arcsec. Abell~401 is to the North-East of the image.}\\label{mosaic} \\end{center} \\end{figure} Numerical simulations (Schindler \\& M\\\"{u}ller 1993, Roettiger, Burns, \\& Loken 1996; Ricker 1998; Takizawa 1999; Ritchie \\& Thomas 2002; Burns et al. 2004) have advanced to such an extent that they can provide us with detailed temperature maps and synthetic X-ray images of clusters under formation, at each stage of their evolution. All this work has demonstrated that cluster mergers result in dramatic substructure, visible in the X-ray images and spectra of clusters of galaxies. Initially, during the first approach of two clusters a compression and/or shock wave is developed in the interface between them. When the two cores collide, the temperature and luminosity of the system reaches its maximum. As the two gravitational potentials settle down to form a single one, two lens-shaped shock waves propagate outwards, along the direction of merging. Soon afterwards, a single remnant is formed. Unfortunately, we have not been able yet to test observationally in a systematic manner, that cluster mergers proceed as theory predicts. Markevitch et al. (1998) set the framework for such investigations, and revealed the gross spectral characteristics of disrupted clusters. Additionally, they were able to comment on the statistics of mergers, but their data were lacking the necessary quality to probe the details of the observed structures. But the results of some recent observational work have been encouraging (e.g., Sun et al. 2002; Kempner, Sarazin, \\& Ricker 2002), as a hotter bar, for example, has been found in a few cases separating the two colliding units, providing the means to calculate the relative velocity of the merging subunits. A detailed comparison of a sample with the numerical simulations is still lacking, and it is unclear whether the simulations include all the necessary physics and are able to model the problem correctly. There are a few examples of cluster properties that have been discovered by the observations, and had not been predicted by the simulations. Recently, for example, X-ray observations have uncovered features (cold fonts; e.g., Markevitch et al. 2000), that have been now explained as signatures of merging clusters. Cold fronts had not been predicted by the simulations, mainly due to the lack of the adequate resolution and the mishandling of the cooling functions. Only recently some numerical simulations have managed to reproduce them and associate their presence to the motion of a galaxy or group through the ICM of another cluster, and explain the observations (Bialek, Evrard, \\& Mohr 2002). Ideally, one would like to construct a sample of clusters that are at different stages of their evolution, and trace the merger event. Unfortunately, finding clusters at different stages of the merging sequence is not trivial. Projection effects or past inaccurate determinations of their physical properties influence severely their classification and taxonomy in one of the stages. Abell~399 and Abell~401 appear at a first sight as a very good example of two clusters at early stages of merging. They are equally rich, and past X-ray missions have found them to be both at temperatures between 7 and 8~keV. They have been the subject of past investigations and have been observed in different wavelengths, from the optical to X-ray. Their projected separation is $\\sim$36~arcmin~$\\sim$~2.96~Mpc\\footnote{Throughout this paper we use $H_0=71~{\\rm km \\ s^{-1} \\ Mpc^{-1}}$, $\\Omega_{\\rm M}$=0.3, and $\\Omega_{\\rm \\Lambda}$=0.7.}, approximately 1-2 times their virial radii. The line-of-sight velocity difference of the two is of the order of $\\sim 700 \\; {\\rm km \\ s^{-2}}$ (Girardi et al. 1997). The velocity dispersions of Abell~399 and 401 are $\\sim 1180$ and $\\sim 1110$~${\\rm km \\ s^{-2}}$ (Oegerle \\& Hill 2001) respectively. Dynamical models of the system based on the dynamics of the galaxies around the Abell~399/401 system reached the conclusion that it consists a bound cluster pair (Oegerle \\& Hill 1994), and that the galaxies of both clusters should move in the total gravitational potential of both clusters. Additionally, since their line-of-sight velocity difference is smaller than their velocity dispersions, it is probable that the encounter between the two clusters is taking place essentially on the plane of the sky. Basic properties of the clusters are given in Table~1. \\begin{table} \\begin{center} \\caption{Target Information}\\label{target_info} \\begin{tabular}{ccccccccc} \\hline \\hline Cluster\t\t& z\t\t& kpc/''\t\t& $D_L$ (Mpc) \\\\ \\hline A399\t\t& 0.0724\t\t& 1.36 \t\t& 322.4 \\\\ A401\t\t& 0.0737\t\t& 1.38\t\t& 328.5 \\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\begin{table*} \\caption{Pointing Information}\\label{obs_info} \\begin{center} \\begin{tabular}{ccccccccc} \\hline \\hline (I)\t\t& (II)\t\t& (III)\t\t& (IV)\t\t& (V)\t\t& (VI)\t\t& (VII) \\\\ Rev \t\t& Obs\t\t& $\\alpha$~(2000) \t& $\\delta$~(2000) \t& Instr.\t\t\t& $Exp$\t\t& $Exp_{corr}$\t\t& \\\\ & & deg \t\t& deg \t\t& & ksec \t\t& ksec\t\t& \\\\ \\hline 0127\t\t& 0112260101\t& 44.4491255188460\t& 13.0518334856848\t& MOS1\t\t\t& 14.166 \t& 11.429 \t& \\\\ & & & & MOS2\t& 14.170\t& 11.599\t& \\\\ & & & & PN\t& 9.188\t& 5.249 \t& \\\\ 0127\t\t& 0112260201\t& 44.5827088537386\t\t& 13.3186112665767\t\t& MOS1\t\t\t\t& 18.231\t\t& 18.133 \t\t& \\\\ & & & & MOS2\t& 18.234\t& 18.152\t& \\\\ & & & & PN\t& 12.505 \t& 12.395 \t& \\\\ 0395\t\t& 0112260301& 44.7636255225170\t& 13.5472223803563 & MOS1\t\t\t& 12.953\t& 12.948\t& \\\\ & & & & MOS2\t& 12.959 \t& 12.897\t& \\\\ & & & & PN\t& 8.086 \t& 8.063\t& \\\\ 0395\t\t& 0112260401\t& 44.9844588584281\t& 13.7638057162178\t& MOS1\t\t\t& 11.913 \t& 11.882\t& \\\\ & & & & MOS2\t& 11.917\t& 11.829\t& \\\\ & & & & PN\t& 7.143\t& 7.143 \t& \\\\ \\hline \\end{tabular} \\vspace{0.2cm} \\begin{minipage}{16cm} \\small NOTES-- (I): revolution number; (II): observation number; (III): pointing $Right Ascension$ ($\\alpha$) in degrees; (IV) pointing $Declination$ ($\\delta$) in degrees; (V) EPIC Instrument; (VI) Exposure time (live time for the central CCD); (VII) Reduced Exposure time, after the subtraction of the bright background flares (see text for more details). \\end{minipage} \\end{center} \\end{table*} \\begin{figure*} \\begin{center} \\setlength{\\unitlength}{1cm} \\begin{picture}(8,7) \\put(-6.5,7.5){\\special{psfile=A399_n.ps angle=-90 hscale=40 vscale=40}} \\put(3,7.5){\\special{psfile=A401_n.ps angle=-90 hscale=40 vscale=40}} \\end{picture} \\end{center} \\caption{\\xmm images of Abell~399 (left panel) and Abell~401 (right panel). The images have been smoothed with Gaussian kernels of $\\sigma$ = 2~pixels=16~arcsec. The four sectors, centred on the central cDs in each cluster, and used in the subsequent analysis, are shown.}\\label{A399pies} \\end{figure*} Previous studies have reached contradictory conclusions about the past history of the pair. The first attempts to reveal signs of interactions in the region between the two with the {\\it Einstein} observatory did not lead to any strong conclusions (Ulmer \\& Cruddace 1981). The \\asca satellite found evidence for some enhancement in the X-ray flux above what is expected from just the superposition of the two clusters (Fujita et al. 1996). A slight temperature increase in-between them was also recorded. Both results led Fujita et al. (1996) to suggest that Abell~399/401 is a pre-merging pair. However, this scenario provides no explanation for the lack of cooling flows in the cores of both clusters. X-ray observations (e.g., Peres et al. 1998) have found that their centres do not host cooling flows as the mass accretion rate is zero. Additionally, the model of Fujita et al. does not tie in well with the past \\rosat observational facts that both clusters appear disrupted : the work of Slezak, Durret, \\& Gerbal (1994), for example, found evidence for `substructure' in Abell~401, and the \\rosat HRI detector (Fabian, Peres, \\& White 1997) revealed a linear structure that emanates from the centre of Abell~399 and points towards Abell~401. This feature lead to the suggestion that the two clusters have already encountered each-other, and are now moving apart. In order to recover the dynamical state of this system, decide on its past history and future evolution, and derive vital information that would help us to test the results of the numerical simulations of merging clusters, we observed Abell~399/401 with \\xmm. The observations are presented in Section~2. Section 3 is devoted to the presentation of the properties of each cluster individually, as found from the \\xmm data and analysis. In Section~4 the X-ray properties of the region between the two is investigated, while in Section~5 the radio properties of the clusters are presented, and their large scale environment is discussed in Section~6. Finally, in the discussion section (Section~6) we present models for the dynamical states of both clusters that can fit well the \\xmm radio and optical results. ", "conclusions": "We summarise here the main results of the \\xmm study of the Abell~399/401 binary cluster system and the conclusions about the its dynamical state that are favoured by them. \\begin{itemize} \\item{The \\xmm data confirm the lack of cooling flows in the cores of both clusters.} \\item{The image analysis gives $\\beta$ values that are lower than the expected canonical value of 0.65 found in rich clusters of galaxies. In neither cluster is the gas azimuthally symmetric around the central cD galaxy.} \\item{A 2-dimensional analysis reveals a lop-sided excess ($<$200~kpc) around the central galaxies.} \\item{There is temperature structure in the inner (200-400)~kpc of both clusters, which argues for the presence of complex structures within the cluster cores, possibly due to shock waves and cold fronts.} \\item{Abell~399 shows a sharp edge to the East of the cluster centre, which is associated with harder X-ray emission. The temperature profile to the North of Abell~401 declines steeper than expected, and there is a lack of flux in the same area.} \\item{Both clusters appear to host central radio halos, although the one seen in the NVSS data of Abell~399 requires confirmation. This halo in Abell~399 appears to follow an edge apparent in the X-ray images, and is associated with harder X-ray emission.} \\item{There is a plethora of tailed radio galaxies in and around both systems.} \\item{We find that the flux from the space between the two clusters is slightly enhanced, above what is expected from the superposition of the two clusters. This region appears also hotter than the region on the opposite side of the cluster core, and the temperature distributions predicted for relaxed clusters by numerical simulations.} \\item{The large scale environment around the binary system appears surprisingly empty. There is no evidence in the galaxy distribution and RASS data for the presence of other clusters nearby.} \\end{itemize} The `substructure' \\xmm finds in the core cluster regions cannot be explained by the tidal field generated by the presence of the other cluster. On larger scales, (10-20)~arcmin$\\simeq$(0.8-1.6)~Mpc from the cluster centres, tidal forces become gradually far more important, in shaping the gas distributions. However, we find that the extensions of the gases towards the other member of the system seen in the \\xmm data cannot again be solely due to their mutual interactions. If the gases were extended by gravitational forces, their temperature should have been lower than what we observe. It is apparent that an additional driving force is required. We also find that the properties of Abell~399/401 cannot be reproduced by scenarios that either involve off-centre cluster collisions, or that speculate that they have been through each other once. An offset collision provides the only single mechanism which might explain all the properties of the system, but that it doesn't seem consistent with the observed large separation and limited disturbance of the clusters at large radii from the cluster centres. On the other hand our analysis finds evidence for increased flux and temperature in the region in-between the two. It appears that the two have already stated interacting mildly, and that we are witnessing a compression region between them. The most possible explanation for the properties we find in the core region of the clusters is that they are due to each ones past merger activity. It seems that each cluster is a merger remnant. A similar evolutionary scenario has been recently proposed by Belsole et al. (2003) to explain the unequal system Abell~1750. Indeed, in a hierarchical structure formation scenario, binary cluster systems might be the late stages of cluster formation, just before the final merger takes place and one rich cluster is formed. The numerical simulations of Frenk et al. (1996) show two clusters forming at a close proximity, by the merging of smaller units, before the two finally merge to form a single rich cluster at present day. Additional support for our proposed model comes from the inspection of simulated X-ray images and temperature maps (from the Simulated X-ray Cluster Data Archive, http://sca.ncsa.uiuc.edu/). These simulations show the X-ray flux and temperature stucture of clusters as they are formed from the continuous accretion of smaller structure. An extensive search though this archive indicates that pairs of large clusters are formed, indeed by the merging of filaments from directions that do not necessarily coinside with the direction to the nearest massive cluster. Additionally, the development of shock waves can be witnessed when the two clusters are at close separations. Additionally, they confirm the presence of features like the indentaion and shock wave to the East of Abell~399, as being produced during the infall of a smaller group. Hence, for this pair of galaxies, we favour a model which combines some early interaction, with a minor merger history." }, "0403/astro-ph0403390_arXiv.txt": { "abstract": "Deep spectropolarimetric observations, obtained with the Very Large Telescope (VLT), are presented for two powerful radio galaxies, 0850--206 (z=1.3373) and 1303+091 (z=1.4093). These observations cover the rest-frame wavelength range $\\sim$\\,1450\\,--\\,3750\\,\\AA. New radio observations and continuum images of the same sources are also presented. These galaxies are the first two observed from a complete sample of nine radio sources with redshifts in the range 1.3\\,$\\le$\\,z\\,$\\le$\\,1.5 (selected from the equatorial sample of powerful radio sources of Best, R{\\\"o}ttgering \\& Lehnert\\nocite{best99b}), as part of a project aimed to investigate the multi-component nature of the UV continuum in radio galaxies and, in particular, any variations of the continuum properties with the radio source age. The larger radio source of the two, 0850--206, presents a high continuum fractional polarization, averaging 17 per cent across the observed wavelength range and reaching 24 per cent at rest-frame wavelengths of $\\lesssim$\\,2000\\,\\AA. The smaller radio source, 1303+091, shows a lower continuum polarization, averaging 8 per cent and rising to 11 per cent for rest-frame wavelengths $\\gtrsim$\\,3000\\,\\AA. For both galaxies, the position angle of the electric vector is generally constant with wavelength and within $\\sim$\\,15$^{\\circ}$ of perpendicular to the radio axis. Both their total flux spectra and polarized flux spectra reveal the 2200\\,\\AA \\ dust feature, and comparison with dust scattering models suggests that the composition of the dust in these galaxies is similar to that of Galactic dust. In 0850--206, scattered quasar radiation dominates the UV continuum emission, with the nebular continuum accounting for no more than $\\sim$\\,22 per cent and no requirement for any additional emission component such as emission from young stars. By contrast, in 1303+091, unpolarized radiation could be a major constituent of the UV continuum emission, with starlight accounting for up to $\\sim$\\,50 per cent and the nebular continuum accounting for $\\sim$\\,11 per cent. The emission-line properties of the galaxies are also studied from their total intensity spectra. Comparison of the measured emission-line ratios with both shock- and photo-ionization models shows that the nuclear and extended gas in these galaxies is mainly photoionized by the central active nucleus. ", "introduction": "The galaxies associated with powerful extragalactic radio sources are uniquely important for understanding the physics of active galactic nuclei (AGN) and for studying the relationship between radio source activity and the properties of the host galaxy and its environment. Powerful distant radio galaxies are thought to be the progenitors of present day, giant ellipticals (e.g. \\pcite{best98,mclure2000}). However, compared with normal elliptical galaxies, powerful radio galaxies at z$\\gtrsim$0.6 show enhanced optical/UV continuum and line emission, which is generally elongated and aligned along the radio axis (e.g. \\pcite{mccarthy87}). The emission-line structures usually extend large distances from the nucleus (5\\,$\\sim$\\,100 kpc; e.g. \\pcite{tadhunter86,baum88}); the properties of these extended emission-line regions (EELR) can be greatly influenced by shocks resulting from interactions between the radio source structures and the interstellar/intergalactic medium (ISM/IGM) (e.g. \\pcite{clark98,carmen01,carmen03}). Interestingly, the emission-line properties of a sample of z$\\sim$1 radio galaxies have been found to evolve strongly as the radio source passes through the host galaxy \\cite{best2000b}: large radio sources ($\\gtrsim$150 kpc) show quiescent `rotation-dominated' velocity profiles and ionization states in agreement with AGN-illumination, while smaller radio sources present highly distorted kinematic profiles and overall ionizations consistent with being shock-dominated. As for the optical/UV continuum excess observed in radio galaxies, several different mechanisms are known to contribute to it (e.g. \\pcite{tadhunter2002}), but the relative extent of their contributions remains uncertain. A popular early interpretation for this UV excess was recent star formation, induced by the passage of the radio jet through~the~ISM of the host galaxy or by merger events linked to the radio source triggering (e.g. \\pcite{rees89}). Jet-induced star formation has been shown to be feasible in numerical simulations (e.g. \\pcite{mellema2002}). Also, stellar absorption features characteristic of OB stars have been observed in\\,the radio\\,galaxy 4C41.17 at z=3.8 (Dey et al.\\,1997)\\nocite{dey97}. The detection of polarized continuum and polarized broad permitted emission lines in some radio galaxies (e.g. \\pcite{antonu-mill85,di-serego89,cimatti93}) suggested that light emitted anisotropically by a hidden quasar nucleus and scattered towards us by dust or electrons in the ISM of the radio galaxy \\cite{tadhunter88,fabian89} also makes an important contribution to the UV light. This is favoured by orientation-based unified schemes for radio sources \\cite{barthel89}, according to which radio galaxies and quasars are drawn from the same parent population but viewed from different angles to the line of sight, with the AGN in radio galaxies obscured by a surrounding dusty torus. Spectropolarimetric studies of a small number of radio galaxies at z$\\sim$1 have shown that the polarized emission is spatially extended, with the electric vector oriented perpendicular to the UV continuum emission at all wavelengths, and that, while the permitted MgII\\,2800 emission line is observed to be broad in polarized light, none of the narrow lines shows significant polarization (e.g. \\pcite{cimatti96}, 1997)\\nocite{cimatti97}. These results clearly show that the origin of the polarization is scattered quasar light. More recent spectropolarimetric studies of a sample of radio galaxies at z$\\sim$2.5 \\cite{vernet2001} have shed light on the nature of the scattering material. These studies show indications of a continuum upturn beyond 2200\\,\\AA \\ in the composite spectrum of the galaxies in the sample, which is interpreted as a possible detection of the 2200\\,\\AA \\ dust feature, indicating that the scattering medium is dust. However, caution must be taken because, due to the high redshift of their sample, those wavelengths are redshifted into a region of strong sky lines. Together with the young stellar and scattered AGN components, other processes known to contribute to the UV continuum excess in some sources are nebular continuum emitted by the extended emission-line gas \\cite{dickson95} and direct AGN light \\cite{shaw95}. As with the emission-line gas, some properties of the continuum emission in radio galaxies are observed to vary with radio source size: smaller sources show strings of bright knots aligned along the radio axis, and larger sources present more compact nebulae with fewer bright components \\cite{best96}. Do the contributions of the different components to the UV excess also evolve over the radio source lifetime? For instance, a large variation is seen in the strengths, and to a lesser extent colours, of the polarized emission in radio galaxies. Is this linked to the radio source size? Such questions cannot be answered by existing spectropolarimetric data, because targets were generally selected on the basis of previously detected polarization in imaging observations, an interesting UV morphology or an ultra-steep radio spectrum, making the samples studied biased or incomplete. To remedy this and address many of the issues raised above, such as the nature of the scatterers at high redshift and the variation with radio source size of the different contributions to the UV excess, we have begun a programme to make deep spectropolarimetric observations of a {\\it complete} sample of nine radio galaxies with redshifts in the range 1.3\\,$\\le$\\,z\\,$\\le$\\,1.5 and a wide range in radio sizes. In this paper we present the observations of the first two galaxies from the sample. The paper is organized as follows. In Section 2, the selection criteria of our sample are described. Section 3 contains the details of the observations, data reduction and analysis. The results are presented in Sections 4 and 5. Discussion of the results follows in Section 6. Summary and conclusions are provided in Section 7. Throughout this paper values of the cosmological parameters of H$_{0}$\\,=\\,65\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_{\\rm M}$\\,=\\,0.3 and $\\Omega_{\\Lambda}$\\,=\\,0.7 are assumed. ", "conclusions": "" }, "0403/astro-ph0403359_arXiv.txt": { "abstract": "Solid observational evidences indicate a strong dependence of the galaxy formation and evolution on the environment. In order to study in particular the interaction between the intracluster medium and the evolution of cluster galaxies, we have created a large database of clusters of galaxies based on the largest available X-ray and optical surveys: the ROSAT All Sky Survey (RASS), and the Sloan Digital Sky Survey (SDSS). We analyzed the correlation between the total optical and the X-ray cluster luminosity. The resulting correlation of $L_X$ and $L_{op}$ shows a logarithmic slope of 0.6, a value close to the self-similar correlation. We analysed also the cluster mass to light ratio, by finding a significant dependence of $M/L$ on the cluster mass with a logarithmic slope ranging from 0.27 in the i and r bands to 0.22 in the z band. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403445_arXiv.txt": { "abstract": "Deep surveys indicate a bubbly structure of cosmological large scale which should be the result of evolution of primordial density perturbations. Several models have been proposed to explain origin and dynamics of such features but, till now, no exhaustive and fully consistent theory has been found. We discuss a model where cosmological black holes, deriving from primordial perturbations, are the seeds for large-scale-structure voids. We give details of dynamics and accretion of the system voids-cosmological black holes from the epochs $(z\\simeq10^{3})$ till now finding that void of $40h^{-1}Mpc$ of diameter and under-density of $-0.9$ will fits the observations without conflicting with the homogeneity and isotropy of cosmic microwave background radiation. ", "introduction": "The existence of voids has been evident after the discovery by Kirshner et al. of a large void with diameter of $60$ Mpc in B\\\"{o}otes (\\cite{kirsh}). Systematic surveys have shown the existence of many regions with similar characteristics. Computer analysis of galaxy distribution gives evidence that voids occupy about 50\\% of the volume of the universe (e.g. see \\cite{El-Ad:1997af}) or, according to a more recent paper (\\cite{hoyle}), about $40$\\% of the volume of the universe. Today, there is a large agreement on the issue that voids are not just empty regions of space, but that they are regions with a very low density of luminous matter. As observed by Peebles (\\cite{peebvoid}), the low dispersion of velocities of galaxies indicates that most of matter must be inside the voids, not only if the density parameter (for the matter component) is $\\Omega_{m}=1$ but also in the case $\\Omega_{m}\\ll 1$. In any case, recent observations (\\cite{boomerang}; \\cite{supernovae1}; \\cite{supernovae2}) indicate that the total value of density parameter is $\\Omega= \\Omega_{m}+ \\Omega_{\\Lambda}=1$ where $\\Omega_{m}\\simeq 0.3$ and $\\Omega_{\\Lambda}\\simeq 0.7$. In this case $\\Omega_{\\Lambda}$ is the contribution due to the whole content of unclustered matter which can be cosmological constant, some kind of scalar field (\\cite{quintessence1}; \\cite{quintessence2}; \\cite{quintessence3}; \\cite{quintessence4}) or, in general, \"dark energy\". It is worthwhile to stress that the visual inspection of galaxy distribution suggests nothing else but the absence of large amount of luminous matter in wide regions. Furthermore, it is not clear whether the voids are spherical regions approximately empty or under-dense regions with arbitrary shapes. Several definitions of voids have been proposed, but a general agreement on their real nature has not been reached yet (\\cite{sch}). The Swiss-Cheese cosmological model, initially proposed by Einstein and Straus (\\cite{einstraus1}; \\cite{einstraus2}), appears suitable for the description of the cosmological voids. In a recent paper (\\cite{cosimo}), it was proposed an approach for the formation of the cosmological voids in the framework of such model. It was shown that voids are the consequence of the collapse of extremely large wavelength perturbations into low-density black holes and of the comoving expansion of matter surrounding the collapsed perturbations. As a result, it was claimed that in the center of each void there is a black hole whose mass $M$ compensates the mass which the void would have if it were completely filled with matter having a cosmological density. In \\cite{cosimo}, the voids are empty regions of the universe which grow comovingly with the cosmological expansion. In that case, the presence of cosmic background radiation was neglected. In this paper, we analyze the physical mechanism capable of explaining the structure of voids in presence of baryonic matter, cosmic background radiation (CBR) with central black holes acting as seeds. The layout of the paper is the following: in Sect.2, we will present the cosmological black hole (CBH) model in the framework of the Friedmann-Lema\\^{i}tre-Robertson-Walker (FLRW) cosmology. Sect.3 is devoted to the discussion of the effects of interaction between the CBR and CBH. A mechanism for the formation of an under-density regime void is analyzed in Sect.4, while the matching with observations, which allows to determine the initial time of voids formation and the mass function of CBHs, is studied in Sect.5. The discussion of results and conclusions are given in Sect.6. ", "conclusions": "In this paper, we have developed a model where cosmological black holes are seeds for large scale structure voids. Such systems come out from the evolution of primordial perturbations and result as stable structures from $(z\\simeq10^{3})$ up to now. They enlarge till diameters of about $40h^{-1}Mpc$ and the under-density of voids is of the order $-0.9$ with respect to the background. The whole structure is a sort of honeycombs where most of galaxies (i.e. luminous matter) are located on the edge of voids while most of dynamical mass is sited in the central black hole. The edge is defined by a natural equilibrium condition on the energy due to the balance of gravitational pull of the central black hole and the cosmic expansion. The cosmic background radiation contributes to the accretion of black hole flowing inside the void but its homogeneity and isotropy is not affected in agreement with data. The picture which emerge agrees with optical and IRAS observation (\\cite{El-Ad:1997af}) giving a $\\sim50\\%$ of the volume filled by voids with the above characteristics. The presence of central black holes seems to confirm the gravitational origin of the voids and stabilizes the system against cosmic expansion preventing its evaporation. It is interesting to note that the order of magnitude observed for the masses of CBH concides with the one of the Great Attractor (\\cite{Fairall}). It is very tempting for us to identify the Great Attractor as a CBH and to use the model described in this paper for explaining the large scale motions observed for the galaxies surrounding it. However, the main problem with observations of cosmic velocity fields (\\cite{Faber}) is that the voids are, in general, the contrary to Great Attractor, and large scale structure around the voids do not show velocity fields converging toward the voids, but toward the visible clusters and superclusters around the voids. This apparent shortcoming, in the framework of our model (see also \\cite{cosimo}), can be overcome by the Birkoff theorem which states that the stationary solutions are also static if the spherical symmetry is restored. So, a fraction of galaxies is attracted by clusters and superclusters \"outside\" the void while another fraction has no dynamics since it has been already attracted \"inside\" the void. This fact could be interpreted as an early selection due to a competitive mechanism between CBHs and external matter contained into clusters and superclusters. However, if the Swiss-Cheese model were always valid such a selection would have never been achieved; instead, in a more realistic situation, the model holds only approximately so then we have to expect galaxies inside and outside the void due to the deviations from sphericity and to the perturbations of the CBH mass. Furthermore, as observed in \\cite{Davis} and \\cite{peebvoid}, the small relative velocity dispersion shows that, if $\\Omega_{M}=1$, then most of the mass has to be contained into the voids. The same authors conjecture that this must be true even when $\\Omega_{M}$ is smaller than 1 as predicted by several CDM simulations." }, "0403/astro-ph0403029_arXiv.txt": { "abstract": "We use the cosmic microwave background angular power spectra to place upper limits on the degree to which global defects may have aided cosmic structure formation. We explore this under the inflationary paradigm, but with the addition of textures resulting from the breaking of a global $O(4)$ symmetry during the early stages of the Universe. As a measure of their contribution, we use the fraction of the temperature power spectrum that is attributed to the defects at a multipole of 10. However, we find a parameter degeneracy enabling a fit to the first-year WMAP data to be made even with a significant defect fraction. This degeneracy involves the baryon fraction and the Hubble constant, plus the normalization and tilt of the primordial power spectrum. Hence, constraints on these cosmological parameters are weakened. Combining the WMAP data with a constraint on the physical baryon fraction from big bang nucleosynthesis calculations and high-redshift deuterium abundance limits the extent of the degeneracy and gives an upper bound on the defect fraction of 0.13 (95\\% confidence). ", "introduction": "\\label{sec:intro} Recent measurements of the cosmic microwave background (CMB), notably the first year WMAP data \\cite{Hinshaw2003, Kogut2003}, have proved highly successful in probing the early stages of cosmic structure formation. The observed CMB anisotropies may be produced by taking adiabatic primordial perturbations of roughly the Harrison-Zel'dovich form and evolving these using well-understood, linear physics. Further, the parameter values that are required for this process to a give a match to the data are consistent with those measured using other astronomical techniques. That the primordial power spectrum predicted by many models of inflation is of the required form has become an important success of the inflationary paradigm. On the other hand, a less attractive property of the paradigm is that successful inflationary models may involve quite different fields, interactions and levels of physical motivation. Here we address the issue using CMB power spectra to constrain models of hybrid inflation \\cite{Linde1994,Lyth1999} that involve the formation of topological defects as inflation ends \\cite{Copeland1994}. Such models, however, do not fit exactly into the above regime. With the existence of topological defects, the seeding of cosmic structure continues after inflation ends, for the defects further perturb the cosmic fluid as long as they continue to be present. For a detailed review of structure formation with defects see Ref. \\cite{Durrer2002}, but generally, a defect-dominated temperature power spectrum does not have pronounced acoustic peaks \\cite{Albrecht1996}. Hence, if defects are added to a passive-evolution case and the normalization reduced to maintain the fit to data on large scales, then the acoustic peaks are slightly suppressed. That is, however, assuming that the other cosmological parameters are not also changed. Earlier work along these lines \\cite{Battye2000, Bouchet2001, Contaldi1999, Jeannerot1997, Pogosian2003} has tended to be of quite a different vein to that described here. In particular we employ a full likelihood analysis for the fit to data in which the cosmological parameters are free to vary. This freedom has important consequences for the independent use of the temperature power spectrum to constrain such hybrid models --- at least given the presently available data. Previous studies have also tended to focus on local defects in the form of cosmic strings, a class of models that we shall not look at here in any detail, and of such work only \\cite{Pogosian2003} has involved data from the WMAP project. In our work, the cosmological perturbations are the uncorrelated sum of those from (i) an inflationary adiabatic model, including both scalar and tensor perturbations, and (ii) a global defect model, with an $O(4)$ symmetry breaking at (or after) the end of inflation producing textures (see e.g. \\cite{Durrer2002}). In parametrizing the primordial perturbations, we take there to be negligible running of the scalar spectral index and that the tensor index obeys the single-field consistency condition (see Sec. \\ref{sec:MCMC}). We also assume negligible neutrino masses, as would result from a hierarchical mass variation with neutrino flavor. The contribution to the CMB power spectra from inflation is found using a variant of the CMBFAST \\cite{Seljak1996} approach, in the form of CAMB \\cite{Lewis2000}. The defect contribution is found by applying the unequal-time correlator (UETC) approach \\cite{Pen1997} to numerical field evolution simulations, as will be discussed in the next section. The UETC method addresses the problem that, in order to calculate CMB power spectra to sub-degree scales, a simulation conventionally requires a dynamic range that is far in excess of that which is feasible with current technology. While it is possible to make full-sky CMB maps \\cite{Pen1994,Landriau2004}, this direct approach is currently limited to relatively low multipoles: $\\ell<20$, although these potentially contain more information than the power spectra alone. The UETC method boosts the dynamic range via a series of theoretical simplifications, for example causality, such that high resolution power spectra calculations may be performed from the simulation data. To do so also involves calculations of the CMBFAST-type, which we carried out using a version of CMBEASY \\cite{Doran2003} modified so as to deal with the defect scenario. Despite the benefits of the UETC approach, the computational requirements of CMB calculations that include defects far exceed those that do not. As a result, the calculation of CMB spectra for a vast number of different cosmological parameter values is not attainable. Hence, the popular Markov chain Monte Carlo (MCMC) approach \\cite{Verde2003, Lewis2002, Gelman1992}, which involves many thousand such calculations, cannot be fully applied to the defect case. However, the non-defect case fits the WMAP data well and defect-dominated structure formation does not give the required acoustic peaks, suggesting that the defect contribution is small. If this is the case, then the result of a small change in the cosmological parameters used for the defect calculation is a second order effect. Therefore, the defect contribution needs only to be calculated once, using currently favored values of the cosmological parameters (see Sec. \\ref{sec:CMBcalc}). The defect contribution is then fixed, except for a normalization factor, which is free since it is not known at which energy scale the defects formed. Hence the approach used here, which is described more fully in Sec. \\ref{sec:MCMC}, is to apply the standard MCMC procedure to the primordial contribution and add in the defect component with its normalization varied as an MCMC parameter. This has been achieved using a slightly modified version of CosmoMC \\cite{Lewis2002}, which is directly linked to CAMB. As this extra parameter controls the degree to which the CMB power spectra differ from the usual non-defect spectra, it shall be the main focus of this paper. However, we shall not present our results in terms of this parameter directly. Rather we shall use the fractional defect contribution to the temperature power spectrum at a particular multipole, $\\ell=10$. The correspondence between the two is roughly linear for low fractions but with a slight spread due to the variation of the non-defect contribution to the chosen multipole. The fractional quantity is, however, more directly understandable. The data that we have used here are principally that from the first year WMAP release \\cite{Hinshaw2003, Kogut2003}: the temperature power spectrum and the temperature-polarization (TE) cross-correlation spectrum. Other CMB projects, such as ACBAR \\cite{Kuo2004}, CBI \\cite{Pearson2003, Readhead2004} and VSA \\cite{Dickinson2004}, which give data out to higher multipoles than WMAP do not provide much in the way of additional constraints on our model. Applying data on cosmological parameters from, for example, work on big bang nucleosynthesis (BBN) \\cite{Kirkman2003} and measurements of the Hubble parameter by the Hubble Key Project (HKP) \\cite{Freedman2001} has proved more important, as will be detailed in section \\ref{sec:MCMC}. The changes to the matter power spectrum that the inclusion of defects causes may be found in an entirely analogous manner to the CMB calculation. However, while there have been recent steps forward in measurements of this from galaxy redshift surveys, such as 2DFGRS \\cite{Percival2001} or SDSS \\cite{Tegmark:2003ud}, and from the Lyman-$\\alpha$ forest \\cite{Croft2002}, we have chosen not to use such data. Galaxy formation in the presence of defect-induced density perturbations is not understood, and even in pure inflation scenarios, inferences from the Lyman-$\\alpha$ forest must be drawn with care \\cite{Seljak:2003jg}. Further, we do not believe that the use of such data would significantly change our results, as we shall discuss in Sec. \\ref{sec:MCMC}. This is a conservative position, driven by our desire to make reliable and statistically meaningful statements about the relative importance of global defects. ", "conclusions": "\\label{sec:con} We have found that in order to constrain the extent to which global defects assisted the seeding of cosmic structure, extra data in addition to the CMB temperature power spectrum must be used, or we require a greater knowledge of the spectrum than we have at present. This is despite the uncertainties in the WMAP data set used being dominated by cosmic-variance over the range where the defect contribution to the power spectrum is greatest. The freedom in the cosmological parameters and the primordial power spectrum are sufficient to allow a significant defect contribution while still fitting the data. It is only when parameters such as the physical baryon density and the Hubble parameter are restricted, via alternative astronomical techniques, that our model is significantly constrained. Then we find that the temperature power spectrum at a multipole of $\\ell=10$ may have a fraction of 0.13 attributed to defects (95\\% confidence). The uncertainty in this upper bound that comes from the MCMC approach is 3\\%. More importantly, the numerical errors in the defect calculation are believed to be of order 10\\%, in which case there may be of order 10\\% change in this result. Also, if the primordial tensor component was removed then this result would increase by about 15\\%. In addition, this bound is quite sensitive to BBN result for the value for the physical baryon fraction $\\Obhh$ used and there has yet to be full agreement in this value among authors. A further result of the degeneracy found is that, if defects were to contribute to cosmic structure formation, then there would a change in the values of the cosmological parameters estimated from the current CMB data. Most notably this affects $\\Obhh$, $h$ and $\\ns$ all of which are subject to an increase upon the addition of defects. This acts to re-enforce the more general point that any inferences made about the cosmological parameters from the WMAP data are model dependent and should be treated with caution. We note that only a single defect type and model has been used in this investigation: textures resulting from the breaking of an $O(4)$ symmetry. However, the contributions from other global defect models are broadly the same as that considered here, although their spectra are by no means identical. Considering the $O(N)$ class of models for $N=2$ (strings), 3 (monopoles), 4 (textures) and 5 (non-topological textures) there is a gradual variation with $N$ of the relative contributions at low multipoles ($\\ell\\sim10$) and high multipoles ($\\ell\\sim300$), with strings giving preference to the latter \\cite{Pen1997}. This is likely to reduce the fractional contribution allowed from strings by perhaps 30\\% at $\\ell=10$. Local defects in the form of cosmic strings may give a broad peak at $\\ell\\sim400$ \\cite{Battye2000, Contaldi1999}, beyond the first acoustic peak, and hence the results may be quite different. Further, we wish to point out that the WMAP project is ongoing and new data with reduced uncertainties, as well as the addition of the EE polarization power spectrum, will be released. While the defect contribution to the EE spectrum is likely to be significant at only at large scales, the freedom for defects when using the CMB power spectra alone will be lessened by these data. It may be, however, that the Planck satellite \\cite{planck} is required to really explore sub-dominant defect contributions using these power spectra independently of other data. This mission will give precise temperature power spectrum measurement at sub-WMAP resolutions. It will also detect the B-mode polarization, which would be produced by the vector and tensor components of the defect perturbations. Finally, the non-Gaussianity of defect-induced perturbations has not been satisfactorily addressed, either in the matter power spectrum or the CMB. This may lead to more sensitive tests of defect scenarios." }, "0403/astro-ph0403265_arXiv.txt": { "abstract": "We present an analysis of high resolution optical spectra for a sample of very young, mid- to late M, low-mass stellar and substellar objects: 11 in the Upper Scorpius association, and 2 (GG Tau Ba and Bb) in the Taurus star-forming region. Effective temperatures and surface gravities are derived from a multi-feature spectral analysis using TiO, \\na and \\pot, through comparison with the latest synthetic spectra. We show that these spectral diagnostics complement each other, removing degeneracies with temperature and gravity in the behavior of each. In combination, they allow us to determine temperature to within 50K and gravity to within 0.25 dex, in very cool young objects. Our high-resolution spectral analysis does not require extinction estimates. Moreover, it yields temperatures and gravities {\\it independent} of theoretical evolutionary models (though our estimates do depend on the synthetic spectral modeling). We find that our gravities for most of the sample agree remarkably well with the isochrone predictions for the likely cluster ages. However, discrepancies appear in our coolest targets: these appear to have significantly lower gravity (by upto 0.75 dex) than our hotter objects, even though our entire sample covers a relatively narrow range in effective temperature ($\\sim$ 300K). This drop in gravity is also implied by inter-comparisons of the data alone, without recourse to synthetic spectra. We consider, and argue against, dust opacity, cool stellar spots or metallicity differences leading to the observed spectral effects; a real decline in gravity is strongly indicated. Such gravity variations are contrary to the predictions of the evolutionary tracks, causing improbably low ages to be inferred from the tracks for our coolest targets. Through a simple consideration of contraction timescales, we quantify the age errors introduced into the tracks through the particular choice of intial conditions, and demonstrate that they can be significant for low-mass objects that are only a few Myr old. However, we also find that these errors appear insufficient to explain the magnitude of the age offsets in our lowest gravity targets. We venture that our results may arise from evolutionary model uncertainties related to accretion, deuterium-burning and/or convection effects. Finally, when combined with photometry and distance information, our technique for deriving surface gravities and effective temperatures provides a way of obtaining masses and radii for susbtellar objects independent of evolutionary models; radius and mass determinations are presented in Paper II. ", "introduction": "In the study of stars, there is perhaps no parameter more fundamental than stellar mass, which is pivotal in determining the entire evolutionary path traced by a star. With the discovery of brown dwarfs, mass determination has become particularly important at the bottom of the Main Sequence. After all, the very notion of brown dwarfs is predicated on mass: these are substellar objects, which is to say they are less massive than the hydrogen-burning limit of $\\sim$ 80 \\mj. The derivation of ultra-low masses has become particularly crucial in the light of new claims of planetary mass objects (which we henceforth contract to ``planemos'') occuring in isolation in star-forming regions (e.g., \\cite{Zapa00, Lucas00}). While both planemos and brown dwarfs are substellar, the distinction between the two is drawn at the fusion boundary of $\\sim$ 13 \\mj: brown dwarfs undergo an initial phase of deuterium fusion, while planemos never harbor any fusion at all. The existence of isolated planemos with a few Jupiter masses, if proven, has significant consequences for both star and planet formation (\\cite{Padoan02, Bate02, Boss01}). However, the key issue of whether the newly discovered free-floating objects really have planetary masses, or are simply misidentified brown dwarfs, remains unsettled for reasons we touch on below. Only a precise mass derivation can unequivocally resolve this question. The empirical determination of substellar masses, though, has so far proven rather difficult. The most direct approach is to apply Kepler's laws to binary (or higher-order) systems with known orbital parameters, in order to obtain a `dynamical mass'. If the components can be directly observed, then the mass can be related to their other properties, such as temperature and luminosity, and the dynamical mass results extended to other directly detected bodies for which a dynamical mass cannot be acquired (e.g., free-floating objects). However, suitable systems with brown dwarf components remain elusive (though a handful are now coming to light; e.g, \\cite{Close02, Potter02, Lane01}). Dynamical masses for planemos in circumstellar planetary systems (i.e., extrasolar planets) {\\it have} been derived, but these objects have not been directly observed, so it is impossible to relate their masses to other physical properties\\footnote{With the exception of the transiting planets HD 209458 and OGLE-TR056 (\\cite{Charbonneau00} and \\cite{Konacki03}, respectively), which have both mass and radius determinations. Also, except in transiting cases, masses from RV surveys are only lower limits, due to the unknown inclination of the system.}. To date, with the exception of HD 209458, no substellar object - whether brown dwarf or planemo - whose intrinsic spectrum has been observed has also been proven to be substellar by a direct, dynamical measurement of its mass. Conversely, no object outside the Solar System with a dynamical substellar mass has actually been directly detected (again with the sole exception of HD 209458)\\footnote{Some of the atmospheric features of the transiting planet HD 209458 have now been directly detected (\\cite{Charbonneau02}; \\cite{Vidal03}). A comparison with the theoretical tracks reveals serious discrepancies between observed and predicted properties, probably due to stellar irradiation effects. A total mass has now also been obtained for one binary system with a spectrum, GJ 569B, indicating at least one brown dwarf component; individual masses should be available in the near future; \\cite{Lane01}.}. Currently, the substellarity of all directly observed bodies is certified either through the lithium test, or by the fact that their effective temperatures are below the minimum Main Sequence temperature (\\cite{Basri00}). However, both tests can only indicate an upper limit to the mass, and not its precise value. Moreover, they are not useful at very early ages, when lithium is undepleted even in low mass stars and substellar objects have not yet cooled sufficiently. For these reasons, precise mass estimates for all directly observed substellar bodies have so far been obtained solely through comparison with theoretical evolutionary tracks. However, the evolutionary models remain largely untested in an absolute sense, and different groups have generated somewhat different tracks (e.g., \\cite{Baraffe98, Burrows97, Dantona94}). Evolutionary modeling uncertainties arise from a variety of sources, such as the treatment of convection, choice of initial conditions and the modifying effects of any initial accretion phase. Additionally, these uncertainties are exacerbated at the earliest ages and for very low masses (\\cite{Baraffe02}). It is precisely for these ages and masses, however, that the evolutionary models are currently most widely used: to distinguish between stellar and substellar objects in young clusters (since the lithium and temperature tests are not robust at these ages), as well as to infer the actual substellar masses. For example, the planemo status of newly identified isolated bodies in young star-forming regions, mentioned earlier, depends entirely on the accuracy of the theoretical tracks. Therefore, given the burgeoning uncertainty in the tracks with decreasing mass and age, it would be extremely useful to have an independent method for determining mass, in order to check the theoretical predictions. We present such a method here, which relies on an accurate measurement of surface gravity from high-resolution spectra. In a nutshell, we first derive surface gravities and effective temperatures (\\teff), by comparing our high-resolution optical spectra to the latest synthetic spectra. We then calculate radii and masses by combining our inferred gravity and \\teff with photometry and distance measurements. In this paper, we present the gravity and temperature spectral analysis. The radius and mass calculations are presented in a companion paper (\\citet{Mohanty03}; hereafter, Paper II). For cool, very low mass objects, the derivation of surface gravities from observed spectra is not trivial. A vast multitude of molecular and atomic spectral lines arise in their low-temperature photospheres; an adequate modeling of these is a prerequite for inferring gravity (and \\teff). This has only become possible in the last several years, due to large advances in computing power and the compilation of increasingly comprehensive linelists. The resulting synthetic spectra are now routinely used to infer the \\teff and model the colors of low-mass stellar and substellar objects (e.g., \\cite{Allard01, Schweitzer01, Leggett01, Leggett00, Burrows00, Basrimoh00, Kirk99}). Here we take the next logical step, of using the synthetic spectra to derive surface gravities as well. \\teff and gravity, together with empirical distance and photometry information, then allow us to calculate masses and radii without resorting to evolutionary model predictions. Clearly, the validity of our analysis depends on that of the synthetic spectra; we will assess the veracity of these spectra at appropriate junctures. It must be pointed out, however, that synthetic spectra are essential even when masses are derived through comparisons with theoretical evolutionary models - either for translating observed spectral types to temperatures (when comparing the data to theoretical HR diagrams), or for converting theoretical temperatures and luminosities to expected photometry (when comparing the data to theoretical color-magnitude diagrams). In other words, both our technique, as well as analyses using evolutionary models, are limited by any shortcomings in the synthetic spectra. In this paper we will derive, through spectral analysis, \\teff and gravities for a sample of young, late-type (i.e., low mass) Pre-Main Sequence objects belonging to the Upper Scorpius and Taurus regions. Since we observe them prior to their main contraction phase, we expect their surface gravities to be lower than in field objects of similar mass. By comparing our results to the predictions of evolutionary models, we will also test these models in the low mass, young age regime where they are most uncertain. ", "conclusions": "The primary conclusion of this paper is that it is quite feasible to derive fairly precise (0.25 dex) gravities in low mass stellar and substellar objects. This can be accomplished with spectra of moderate S/N and high resolution (which generally require 8-m class telescopes), along with the most advanced model atmosphere and spectral calculations. If one performs this for objects whose radii can be found (this generally requires reasonably precise distances and photometry), then masses can be found well enough to distinguish planemos, brown dwarfs, and stars (see Paper II). Along with the gravity, we obtain a good effective temperature ($\\pm$50K), which facilitates the conversion of a luminosity to a radius. The basic reason for our success is that TiO is a spectral diagnostic that is strongly dependent on temperature, and weakly dependent on gravity (and with the opposite gravity dependence of our atomic line diagnostics). Our gravity diagnostics are both subordinate and resonance neutral alkali lines (of potassium and sodium). They have both a temperature and gravity dependence, and yield degenerate solutions in the two parameters. We therefore rely on TiO to break the temperature degeneracy, and the alkali lines then provide a gravity. We note that there are a number of other alkali lines that we have not yet employed (other lines of potassium and sodium, plus lines of rubidium, cesium, and lithium), which cover a range of wavelengths and have different strengths for different stellar temperatures. In the L dwarfs, a substitute for the TiO temperature diagnostic must be found (since TiO has condensed into dust). The metal hydride molecules (FeH and CrH) are the obvious candidates. Properly developed, our methodology should work well over the range of substellar objects from mid-M through L (which encompass a range of masses from the planetary to stellar domains, depending on age). Indeed, once theoretical tracks have been properly computed and calibrated, one might be able to directly infer age as well (in principle). Having applied our methodology to very young, low-mass objects in the Upper Scorpius and Taurus star-forming regions, we find {\\it (1)} a good agreement between our gravities and the theoretical ones for most of our sample, but {\\it (2)} a significant discrepancy between the two for our coolest targets. We show that even without detailed comparison against model spectra, the data themselves suggest a sharp fall-off in gravity in the coolest objects. We have examined various processes and synthetic spectral uncertainties which might lead us to infer erroneous gravities (e.g., dust, cool spots, metallicity variations, and inadequacies in the model treatment of collisional broadening), and found that they are unlikely to give rise to the spectral effects we see: real gravity variations are indicated. Theoretical tracks would interpret this range of gravities as a range of ages, and we provide several arguments against that interpretation. The alternative is that the theoretical tracks are allowing the very low-mass objects to contract too quickly; the gravities we measure for them are lower than predicted. We discuss various ways in which this could happen, and suggest remaining uncertainties in the model treatment of accretion, deuterium-burning and/or convection are most likely responsible. This means that the interpretation of masses and ages from isochrone analysis in young clusters is problematic at the moment for very low-mass bodies. Nonetheless, we show in Paper II that the interpretation of the faintest of these objects as low-mass brown dwarfs or planetary mass objects looks correct. Finally, we point out that our spectral analysis is carried out in the optical, but it could potentially be accomplished at other wavelengths as well. In particular, the near-IR would be very interesting to explore. Such investigations have already begun: \\cite{Doppmann03a} and \\cite{Doppmann03b} present a similar high-resolution study in the near-IR for higher mass, hotter PMS objects, while \\cite{Gorlova03} examine near-IR PMS substellar spectra, albeit at low-resolution. A high-resolution analysis of cool, very low-mass PMS objects would be extremely fruitful, especially for probing the youngest, most extincted bodies; we have recently embarked on a project to accomplish this." }, "0403/gr-qc0403014_arXiv.txt": { "abstract": "We study the anisotropies of the Galactic confusion noise background and its effects on LISA data analysis. LISA has two data streams of the gravitational waves signals relevant for low frequency regime. Due to the anisotropies of the background, the matrix for their confusion noises has off-diagonal components and depends strongly on the orientation of the detector plane. We find that the sky-averaged confusion noise level $\\sqrt {S(f)}$ could change by a factor of 2 in three months, and would be minimum when the orbital position of LISA is either around the spring or autumn equinox. ", "introduction": "The Laser Interferometer Space Antenna (LISA) is planned to be launched around 2011 and is expected to establish a new window in the low frequency gravitational wave astronomy from $0.1{\\rm mHz}$ to $ 100$mHz \\cite{lisa,Cutler:2002me}. Its main astrophysical targets are Galactic binaries and cosmological massive black holes (MBHs) that merge with other MBHs or capture compact objects. As for the Galactic binaries, there would be a lot of sources in the LISA band. For example we will be able to resolve several thousand close white dwarf binaries at $f\\gsim 3$mHz \\cite{bcn1,bcn2}. At lower frequency regime $f\\lsim 3$mHz they are highly overlapped in the frequency bins, and it would be difficult to resolve them individually. As a result, they form a confusion noise background whose magnitude could be larger than the detector noise at $0.1{\\rm mHz}\\lsim f \\lsim 3$mHz \\cite{lisa}. The coalescence frequency $f_c$ of a MBH system is given by its redshifted total mass $M_z$ as $f_c\\sim 2(M_z/2\\times 10^6M_\\odot)^{-1}$mHz. The mass function of MBHs is highly uncertain at the lower end $\\sim 10^5M_\\odot$ \\cite{Ferrarese:2000se}, and we might have to search treasurable signals from cosmological MBHs in the Galactic confusion noise background. The spatial distribution of the Galactic binaries would trace the Galactic structure well, and the confusion noise background would be strongly anisotropic. The background itself can be regarded as a signal that would provide some information on our Galaxy \\cite{gia,Cornish:2001hg,Ungarelli:2001xu}. But we should notice that there exists a more straightforward approach to probe the Galactic structure using thousands of resolved binaries whose three dimensional positions are estimated in some error boxes \\cite{gia}. In this paper we will focus on the role of the background as a noise, and study the effects of its anisotropies on the signal analysis of LISA. We do not pay attention to the normalization of the background that has been studied by other papers using an isotropic approximation of the source distribution \\cite{bcn1}. This paper is organized as follows. In Section 2 we study the response of interferometers to the anisotropic gravitational wave background with using the long wave approximation, and formulate the noise matrix for the Galactic confusion background. In Section 3 we numerically evaluate the bakground noise as a function of the orientation of the detector, and estimate the annual modulation of the noise level for LISA. In Section 4 we discuss the anisotropies of the signal to noise ratio (SNR) of sources at a fixed distance. Section 5 is devoted to a brief summary of this paper. In the Appendix the confusion noise matrix is analyzed in a different manner from Section 2. ", "conclusions": "In this paper we have studied the anisotropies of the Galactic confusion noise background and its effect on the data analysis of LISA at the low frequency regime $f\\lsim 3$mHz. In contrast to the traditional monopole approximation the anisotropies induce the correlation of confusion noises between the $A$ and $E$ modes of LISA, and the effective noise level depends strongly on the time. In the followings we briefly summarize our results. The effective noise level of LISA becomes smallest around the autumn equinox, and largest in summer for the configuration in figure 2. The difference of the noise $\\sqrt{S^B_{eff}}$ at these two epochs is a factor of 2. The detector plane of LISA at the autumn equinox is almost normal to the Galactic poles and the effective noise level is very close to the global minimum of the noise map given for the all possible orientations of the detector. The time average of the effective noise is different from the traditional monopole approximation by less than $50\\%$. We have also analyzed the dependence of the SNR on the sky position of sources with a fixed distance and averaged orientation. For a long lived source the dependence after 1yr integration is weak, and fluctuations on the sky are $\\sim 30\\%$ level. This is mainly due to the angular average of the response function by annual rotation of LISA. For a short lived source the SNR at a given direction changes strongly with time. For example the SNR of a source at the direction of the Galactic pole changes by a factor of $\\sim4 $ in three months. A factor of 2 comes from the modulation of the confusion noise level and another factor of 2 is from that of the angular response function of the detector." }, "0403/astro-ph0403579_arXiv.txt": { "abstract": " ", "introduction": "In this contribution, I review the basic ingredients of the so-called Randall-Sundrum brane cosmology. This new cosmological scenario is based on the assumption that our universe is a {\\it brane}: a sub-space embedded in a bulk spacetime, with a single extra dimension. In contrast with other braneworld models, the self-gravity of the brane is taken into account. As will be recalled in this contribution, the cosmology for such a brane-universe is modified in two respects: \\begin{itemize} \\item the Friedmann equation is modified at high energy; \\item the bulk influences the cosmological evolution via an addititional term, usually called dark radiation or Weyl radiation. \\end{itemize} Whereas the first modification has an impact only during the very early universe, since it is significant only at high energy, the second effect could have observable consequences today, as discussed below. In this contribution, I do not discuss other important topics in the context of brane cosmology, such as the issue of cosmological perturbations, since this topic will be discussed by Roy Maartens. Finally, for the reader who wants to learn more on this subject, he/she can find in the literature several detailed reviews \\cite{reviews} that cover much more than the present contribution. ", "conclusions": "The various models of extra dimensions with branes have raised a considerable interest in the last few years, motivated by their more or less direct connections with the recent developments in string/M theory. To make further progress in this direction, it is important to see how these models can be tested by experiments. Roughly speaking, the tests can be classified into three broad categories: modification of Newton's law; signatures in colliders; cosmology. As usual in high energy physics, if the scale characterizing new physics is too high then it cannot be reached directly in collider experiments. In this case cosmology is the only place where the effects of new physics can be, indirectly, observed. As illustrated by numerous works in the last few years, Randall-Sundrum type cosmology is a very rich playground to study the very peculiar consequences of the braneworld idea in cosmology. As recalled in this contribution, its essential new features are: a $\\rho^2$ term in the generalized Friedmann equation, which dominates at high energies; ``dark radiation'', {\\it produced} during the high energy phase, and of potential relevance for observations, via the nucleosynthesis constraints on the number of extra relativistic degrees of freedom." }, "0403/astro-ph0403053_arXiv.txt": { "abstract": "We present the results of qualitative consideration of possible \\q changes occurring during the transition from the hot accretion disc to the cool one. We argue the possible existence of one more type of spiral density waves in the inner part of the disc where gasdynamical perturbations are negligible. The mechanism of formation of such a wave as well as its parameters are considered. We also present the results of 3D gasdynamical simulation of cool accretion discs. These results confirm the hypothesis of possible formation of the spiral wave of a new, ``precessional\" type in the inner regions of the disc. Possible observational manifestations of this wave are discussed. ", "introduction": "The analysis of principal processes of matter heating and cooling in accretion discs presented in the work [\\ref{di12}] has shown that for realistic parameters of accretion discs in semidetached binaries ($\\mdot\\simeq10^{-12}\\div10^{-7}M_\\odot/\\mbox{year}$ and $\\alpha\\simeq10^{-1}\\div10^{-2}$)\\footnote{$\\mdot$ -- mass transfer rate; $\\alpha$ -- dimensionless parameter introduced by Shakura and Sunyaev [\\ref{shakura}, \\ref{shak2}] for expression of viscosity coefficient $\\nu=\\alpha c_s H$ ($H$ -- disc semithickness, $c_s$ -- sound speed).} the gas temperature in outer part of the disc lies in the range from $\\sim 10^4$~K to $\\sim 10^6$~K. Earlier we have conducted 3D gasdynamical simulation of accretion discs both for a `hot' case (the gas temperature in the outer part of the disc was 200--500 thousands~K [\\ref{di1}--\\ref{di11}]), and for the `cool' case (the temperature of gas in the outer part of the disc didn't exceed $\\sim 2 \\times 10^4$~K [\\ref{di12}]). The analysis of these results has shown that for both cases (i.e. independently on the disc temperature) the self-consistent solution didn't involve the shock interaction between the stream of matter from the inner Lagrangian point $L_1$ and formed accretion disc (``hot spot\"). Energy release zone (``hot line\") is located outside the disc and is formed as the result of the interaction of the circumdisc halo and circumbinary envelope with the stream. The ``hot line\" model is found to be in a good agreement with observations [\\ref{Tanya98}--\\ref{Tanya2003b}]. For `hot' solutions we have investigated both general morphology of gaseous flows in semidetached binaries and the structure of formed hot accretion discs (see, e.g., [\\ref{di11},\\ref{ippeg}]). In particular, we have found only one arm of spiral shock wave generated by the tidal influence of the mass-losing star. The two-armed spiral shocks were discovered by Matsuda {\\it et al.} in [\\ref{Sawada86}--\\ref{spiral2}]. Nevertheless, our 3D gasdynamical simulations for the `hot' case have shown the presence of only one-armed spiral shock while in the place where the second arm should be the stream from $L_1$ dominates and presumably prevents the formation of the second arm of tidally induced spiral shock. Besides we have found out that for the `hot' case the variation of mass transfer rate leads to the disc perturbations and to the formation of spiral-vortex structure in the disc [\\ref{blob1}--\\ref{fridman}]. Even a glance at the morphology of gaseous flows for the `cool' case (see, e.g., [\\ref{di12},\\ref{rujichka}]) discovers that the accretion disc for this case is characterized by principally different parameters as compared to the `hot' case. In particular, the disc has more circular form and the second arm of the tidal spiral wave is present. Our analysis [\\ref{di12}] shows that opposite to the `hot' case tidal spiral waves don't propagate to the inner part of the cool accretion disc and are located in the outer part of the disc only. The aim of this work is the investigation of the structure of cool accretion discs in semidetached binaries. Section~2 contains the qualitative analysis of changes occurring in transition from a hot accretion disc to the cool one. In particular, here we suggest the possible existence of one more type of spiral density waves in the inner part of the disc where gasdynamical perturbations are negligible. We also consider the mechanism of formation of such a wave and its parameters. Section~3 contains the results of 3D gasdynamical simulation of flow structure for the case when radiative cooling is effective and the gas temperature drops down to $\\sim 10^4$~K in the whole region. These results confirm the hypothesis of possible formation of a spiral wave of a new type in the inner regions of the disc. Possible observational manifestations of the discovered spiral wave of the new, ``precessional\" type as well as main conclusions~are~drawn~in~Section~4. \\renewcommand{\\thefigure}{1} \\begin{figure}[t] \\centerline{\\hbox{\\psfig{figure=ris1.eps,width=\\textwidth}}} \\caption{\\small The sketch of main peculiarities of the morphology of gaseous flows in semidetached binaries for the case of high gas temperature.} \\end{figure} \\renewcommand{\\thefigure}{2} \\begin{figure}[t] \\centerline{\\hbox{\\psfig{figure=ris2.eps,width=\\textwidth}}} \\caption{\\small The sketch of main peculiarities of the morphology of gaseous flows in semidetached binaries for the case of low gas temperature.} \\end{figure} ", "conclusions": "The qualitative analysis of possible changes occurring in transition from a hot accretion disc to the cool one shows the possible generation of spiral wave of a new type in the accretion disc. The appearance of this spiral wave in the inner part of the disc where gasdynamical perturbations are negligible is due to retrograde precession of flowlines in the binary system. The analysis of presented results of 3D gasdynamical simulation fully confirms our hypothesis on the possible generation of spiral wave in the inner part of the cool disc. The correspondence between the qualitative analysis and the computational results permits us to argue the precessional mechanism of the wave formation. Increasing of the radial flux of matter after passing the density wave results in growth of accretion rate and formation of a compact zone of energy release on the accretor surface. This zone can be seen as a periodic increase of brightness on light curves of semidetached binaries. Observation of this zone will permit to determine the precession rate of the wave so these observations will both give the proof of the existence of ``precessional\" wave in the inner part of the cool accretion disc and provide the information on characteristics of the inner parts of the disc." }, "0403/astro-ph0403609_arXiv.txt": { "abstract": "Infrared spectroscopy provides a direct handle on the composition and structure of interstellar dust. We have studied the dust along the line-of-sight towards the Galactic Center using Short Wavelength Spectrometer (SWS) data obtained with the Infrared Space Observatory (ISO). We focussed on the wavelength region from 8--13 $\\mu$m which is dominated by the strong silicate absorption feature. Using the absorption profiles observed towards Galactic Center Sources (GCS) 3 and 4, which are C-rich Wolf-Rayet Stars, { as reference objects,} we are able to disentangle the interstellar silicate absorption and the silicate emission intrinsic to the source, toward Sgr A$^*$ and derive a very accurate profile for the intrinsic 9.7 $\\mu$m band. The interstellar absorption band is smooth and featureless and is well reproduced using a mixture of 15.1\\% amorphous pyroxene and 84.9\\% of amorphous olivine { by mass}, all in spherical sub-micron-sized grains. There is no direct evidence for substructure due to interstellar crystalline silicates. By { minimizing} $\\chi^2$ { of spectral fits to the absorption feature} we are able to determine an upper limit to { the degree of crystallinity of silicates} in the diffuse interstellar medium (ISM), and conclude that the { crystalline fraction of the interstellar silicates} is 0.2\\% $\\pm$ 0.2\\% by mass. { This} is much lower than the degree of crystallinity observed in silicates { in the circumstellar environment of} evolved stars, { the main contributors of dust to the ISM}. { There are two possible explanations for this discrepancy.} First, an amorphization process occurs in the { ISM} on a time scale significantly shorter than the destruction time scale, possibly caused by { particle bombardment by} heavyweight ions. Second, we consider the possibility that the crystalline silicates in stellar ejecta are diluted by an additional source of amorphous silicates, in particular supernovae. { We also compare our results with a study on silicate pre-solar grains found in interplanetary dust particles.} ", "introduction": "In the last decade, a multitude of evidence for the presence of crystalline silicates in various astrophysical environments has emerged. In particular, infrared spectra have revealed that silicates in circumstellar environments often contain a significant crystalline fraction, both around post-main-sequence stars \\citep[e.g.][]{WMJ_96_mineralogy,MWT_02_xsilI} and pre-main-sequence stars \\citep[e.g.][]{WWD_96_xsilsyoung,MWB_01_haebe}. In addition, crystalline silicates are ubiquitous in the Solar System, not only in the more evolved bodies such as planets, but also in primitive objects like comets \\citep[e.g.][]{W_02_cometgrains}. Because crystallization is inhibited by high energy barriers, the origin and evolution of the crystalline silicate fraction in interstellar and circumstellar media has the potential to provide direct evidence of the energetic processing of grains. The life cycle of dust starts in the outflow of evolved stars, continues upon ejection in the interstellar medium and eventually ends in the planet forming disk around a young star. It is surprising that, whereas at the beginning and end of dust grains' lifes crystallinity is prevalent, no crystallinity is found in the intermediate phase (i.e.~in the diffuse interstellar medium). In fact, a relatively high upper limit to the degree of crystallinity in the diffuse ISM has been determined recently \\citep{LD_01_silicate}. Less than 5\\% by number of the interstellar Si-atoms were found to be incorporated in crystalline silicate grains of $<1$ $\\mu$m in size, which is roughly equivalent to a mass fraction of $< 5$\\%. On the other hand, \\citet{BA_02_mineralogy} have suggested that, while the broad and structureless interstellar 10 $\\mu$m absorption feature is commonly ascribed to amorphous silicates, the crystalline spectral detail may be washed out in a very complex mixture of crystalline silicates. Furthermore, some studies on silicates in the dense ISM have reported the (controversial) detection of crystalline silicate features \\citep{CJL_00_xsil_in_orion,OO_03_carbononions}, but one has to bear in mind that this environment has very different physical properties than the diffuse ISM. In this work, we will re-address the issue of crystallinity in the diffuse ISM by studying the line-of-sight towards the Galactic Center. Because of its large amount of extinction and its high infrared flux, the sightline towards the Galactic Center has often been used to characterize the properties of interstellar dust \\citep{RA_85_extinction,RRP_89_gc,PSA_94_GC,TWA_96_GC,LFG_96_gc,CPG_98_hydrocarbon}. We will study the 10 $\\mu$m silicate absorption feature in order to determine the degree of crystallinity in this line-of-sight. In Sect.~\\ref{sec:obs} we will discuss the ISO SWS observations and data reduction, as well as the characteristics of the region around Sgr A$^{*}$ and the correction method applied for emission intrinsic to the Galactic Center (GC) region. The method used to determine the dust composition in the diffuse ISM is described in Sect.~\\ref{sec:determination} along with the results. In Sect.~\\ref{sec:processing} we discuss two mechanisms to explain the discrepancy in crystallinity observed between stellar ejecta and the diffuse ISM. A comparison with silicates in the solar system and planet forming disks around other stars is given in Sect.~\\ref{sec:comparison}. Sect.~\\ref{sec:conclusions} contains the conclusions. ", "conclusions": "\\label{sec:conclusions} In this work, the results of a study on the crystallinity of silicates in the diffuse ISM have been presented. It is known that crystalline silicates are common in the circumstellar environment of both pre- and post-main-sequence stars. It is generally accepted that the dust particles found in the circumstellar environment of young stars originate from evolved stars, and have arrived in their current location after a long (several Gyr) residence time in the interstellar medium. The puzzling lack of evidence of crystalline silicates in the ISM, prompted us to set a firm upper limit on the crystallinity. Studying the 10 $\\mu$m silicate feature in absorption towards Sgr A$^*$ we are able to determine that at most 0.4\\% of the silicates in the interstellar medium have a crystalline structure. The data are best fitted with a degree of crystallinity of 0.2\\%. In addition, we have determined the composition of the amorphous silicate component in the diffuse interstellar medium. We found that 84.9\\% of the amorphous grains are olivine (Mg$_{2x}$Fe$_{2(1-x)}$SiO$_4$) and 15.1\\% are pyroxene (Mg$_{x}$Fe$_{(1-x)}$SiO$_3$). The amorphous grains were found to be spherical and smaller than $\\sim 0.1$ $\\mu$m in radius. { Detailed analysis of the 10 $\\mu$m feature indicates that the pyroxenes are probably slightly Mg-rich (with $0.5 < x < 0.6$) and the interstellar olivines may be Fe-rich (with $0.4 < x < 0.5$). } By comparison of these results with the crystallinity of the silicates produced by mass-losing stars, and the interstellar grain destruction rate, we have determined the interstellar amorphization rate, and found that crystalline silicates are effectively amorphitized in 5 Myr to achieve a final crystallinity of 0.2\\%, while an amorphization time scale of 9 Myr is consistent with the determined upper limit of 0.4\\%. These numbers are only estimates, as the crystallinity of stellar ejecta is not determined very accurately yet. Amorphization by low energy { ion bombardment} has been explored as an explanation for the low degree of crystallinity of silicates in the ISM, but the results { are not yet} consistent with the observed low degree of crystallinity. In addition, dilution by other sources of amorphous silicates, such as supernovae, are not sufficient to explain the observed lack of crystalline silicates in the ISM. We conclude that effectively all silicates in the diffuse interstellar medium become amorphous on a very short time scale, compared to the total residence time in the diffuse ISM. Hence, the crystalline silicates found in the circumstellar environments of young stars, in the solar system and in star formation regions have not survived the ISM unaltered, but are probably crystallized locally. Both annealing as well as evaporation and subsequent condensation seem to be significant crystallization processes." }, "0403/astro-ph0403323_arXiv.txt": { "abstract": "We present here a scenario, based on a low reheating temperature $ T_R \\ll 100$~MeV at the end of (the last episode of) inflation, in which the coupling of sterile neutrinos to active neutrinos can be as large as experimental bounds permit (thus making this neutrino ``visible'' in future experiments). In previous models this coupling was forced to be very small to prevent a cosmological overabundance of sterile neutrinos. Here the abundance depends on how low the reheating temperature is. For example, the sterile neutrino required by the LSND result does not have any cosmological problem within our scenario. ", "introduction": " ", "conclusions": "" }, "0403/nucl-th0403084_arXiv.txt": { "abstract": "Cross sections for the photon-induced particle-emission reactions ($\\gamma$,n), ($\\gamma$,p), and ($\\gamma$,$\\alpha$) are given for all natural isotopes from Ti to Bi. The target nuclei are assumed to be in their ground states, except for $^{180}$Ta which is naturally occurring as the isomer $^{180\\mathrm{m}}$Ta. The cross sections are calculated in a statistical model (Hauser-Feshbach) approach and covering an energy range from threshold up to 7.35 MeV above the threshold (14.7 MeV above threshold for ($\\gamma$,$\\alpha$) reactions). The results are intended to aid conception and analysis of experiments which can also be used to test the methods involved in predicting astrophysical reaction rates for nucleosynthesis. ", "introduction": " ", "conclusions": "" }, "0403/nucl-th0403059_arXiv.txt": { "abstract": "The possibility of appearance of spin polarized states in symmetric and strongly asymmetric nuclear matter is analyzed within the framework of a Fermi liquid theory with the Skyrme effective interaction. The zero temperature dependence of the neutron and proton spin polarization parameters as functions of density is found for SkM$^*$, SGII (symmetric case) and SLy4, SLy5 (strongly asymmetric case) effective forces. By comparing free energy densities, it is shown that in symmetric nuclear matter ferromagnetic spin state (parallel orientation of neutron and proton spins) is more preferable than antiferromagnetic one (antiparallel orientation of spins). Strongly asymmetric nuclear matter undergoes at some critical density a phase transition to the state with the oppositely directed spins of neutrons and protons while the state with the same direction of spins does not appear. In comparison with neutron matter, even small admixture of protons strongly decreases the threshold density of spin instability. It is clarified that protons become totally polarized within a very narrow density domain while the density profile of the neutron spin polarization parameter is characterized by the appearance of long tails near the transition density. ", "introduction": "The spontaneous appearance of spin polarized states in nuclear matter is the topic of a great current interest due to relevance in astrophysics. In particular, the effects of spin correlations in the medium strongly influence the neutrino cross section and neutrino luminosity. Hence, depending on whether nuclear matter is spin polarized or not, drastically different scenarios of supernova explosion and cooling of neutron stars can be realized. Another aspect relates to pulsars, which are considered to be rapidly rotating neutron stars, surrounded by strong magnetic field. There is still no general consensus regarding the mechanism to generate such strong magnetic field of a neutron star. One of the hypotheses is that magnetic field can be produced by a spontaneous ordering of spins in the dense stellar core. The possibility of a phase transition of normal nuclear matter to the ferromagnetic state was studied by many authors. In the gas model of hard spheres, neutron matter becomes ferromagnetic at $\\varrho\\approx0.41\\,\\mbox{fm}^{-3}$~\\cite{R}. It was found in Refs.~\\cite{S,O} that the inclusion of long--range attraction significantly increases the ferromagnetic transition density (e.g., up to $\\varrho\\approx2.3\\,\\mbox{fm}^{-3}$ in the Brueckner theory with a simple central potential and hard core only for singlet spin states~\\cite{O}). By determining magnetic susceptibility with Skyrme effective forces, it was shown in Ref.~\\cite{VNB} that the ferromagnetic transition occurs at $\\varrho\\approx0.18$--$0.26\\,\\mbox{fm}^{-3}$. The Fermi liquid criterion for the ferromagnetic instability in neutron matter with the Skyrme interaction is reached at $\\varrho\\approx2$--$4\\varrho_0$~\\cite{RPLP}, where $\\varrho_0=0.16\\,\\mbox{fm}^{-3}$ is the nuclear matter saturation density. The general conditions on the parameters of neutron--neutron interaction, which result in a magnetically ordered state of neutron matter, were formulated in Ref.~\\cite{ALP}. Spin correlations in dense neutron matter were studied within the relativistic Dirac--Hartree--Fock approach with the effective nucleon--meson Lagrangian in Ref.~\\cite{MNQN}, predicting the ferromagnetic transition at several times nuclear matter saturation density. The importance of the Fock exchange term in the relativistic mean--field approach for the occurrence of ferromagnetism in nuclear matter was established in Ref.~\\cite{TT}. The stability of strongly asymmetric nuclear matter with respect to spin fluctuations was investigated in Ref.~\\cite{KW}, where it was shown that the system with localized protons can develop a spontaneous polarization, if the neutron--proton spin interaction exceeds some threshold value. This conclusion was confirmed also by calculations within the relativistic Dirac--Hartree--Fock approach to strongly asymmetric nuclear matter~\\cite{BMNQ}. For the models with realistic nucleon--nucleon (NN) interaction, the ferromagnetic phase transition seems to be suppressed up to densities well above $\\varrho_0$~\\cite{PGS}--\\cite{H}. In particular, no evidence of ferromagnetic instability has been found in recent studies of neutron matter~\\cite{VPR} and asymmetric nuclear matter~\\cite{VB} within the Brueckner--Hartree--Fock approximation with realistic Nijmegen II, Reid93 and Nijmegen NSC97e NN interactions. The same conclusion was obtained in Ref.~\\cite{FSS}, where magnetic susceptibility of neutron matter was calculated with the use of the Argonne $v_{18}$ two--body potential and Urbana IX three--body potential. Here we continue the study of spin polarizability of nuclear matter with the use of an effective NN interaction. As a framework of consideration, a Fermi liquid (FL) description of nuclear matter is chosen~\\cite{AKP,AIP}. As a potential of NN interaction, we use the Skyrme effective interaction, utilized earlier in a number of contexts for nuclear matter calculations~\\cite{SYK}--\\cite{AAI}. We explore the possibility of FM and AFM phase transitions in nuclear matter, when the spins of protons and neutrons are aligned in the same direction or in the opposite direction, respectively. In contrast to the approach, based on the calculation of magnetic susceptibility, we obtain the self--consistent equations for the FM and AFM spin order parameters and find their solutions at zero temperature. This allows us to determine not only the critical density of instability with respect to spin fluctuations, but also to establish the density dependence of the order parameters and to clarify the question of thermodynamic stability of various phases. The main emphasis in our study will be laid on the region of zero isospin asymmetry (symmetric nuclear matter) and large isospin asymmetry (strongly asymmetric nuclear matter and neutron matter). Note that we consider the thermodynamic properties of spin polarized states in nuclear matter up to the high density region relevant for astrophysics. Nevertheless, we take into account the nucleon degrees of freedom only, although other degrees of freedom, such as pions, hyperons, kaons, or quarks could be important at such high densities. ", "conclusions": "Spin instability is a common feature, associated with a large class of Skyrme models, but is not realized in more microscopic calculations. The Skyrme interaction has been successful in describing nuclei and their excited states. In addition, various authors have exploited its applicability to describe bulk matter at densities of relevance to neutron stars \\cite{SMK}. The force parameters are determined empirically by calculating the ground state in the Hartree--Fock approximation and by fitting the observed ground state properties of nuclei and nuclear matter at the saturation density. In particular, the interaction parameters, describing spin--spin and spin--isospin correlations, are constrained from the data on isoscalar \\cite{T,LS} and isovector (giant Gamow--Teller)~\\cite{HHR,SGE,BDE} spin--flip resonances. In a microscopic approach, one starts with the bare interaction and obtains an effective particle--hole interaction by solving iteratively the Bethe--Goldstone equation. In contrast to the Skyrme models, calculations with realistic NN potentials predict more repulsive total energy per particle for a polarized state comparing to the nonpolarized one for all relevant densities, and, hence, give no indication of a phase transition to a spin ordered state. It must be emphasized that the interaction in the spin-- and isospin--dependent channels is a crucial ingredient in calculating spin properties of isospin symmetric and asymmetric nuclear matter and different behavior at high densities of the interaction amplitudes, describing spin--spin and spin--isospin correlations, lays behind this divergence in calculations with the effective and realistic potentials. In this study as a potential of NN interaction we chose SkM$^*$ and SGII (symmetric nuclear matter) as well as SLy4 and SLy5 (strongly asymmetric nuclear matter) Skyrme effective forces. The models SkM$^*$ and SGII~\\cite{SG} have been constrained by fitting the properties of nucleon systems with very small isospin asymmetries, while the models SLy4 and SLy5 were further constrained to reproduce the results of microscopic neutron matter calculations (pressure versus density curve)~\\cite{CBH}. Besides, in a recent publication~\\cite{SMK} it was shown that the density dependence of the nuclear symmetry energy, calculated up to densities $\\varrho\\lesssim3\\varrho_0$ with SLy4 and SLy5 effective forces (together with some other sets of parameters among the total 87 Skyrme force parametrizations checked) gives the neutron star models in a broad agreement with the observables, such as the minimum rotation period, gravitational mass--radius relation, the binding energy, released in supernova collapse, etc. This is an important check for using these parametrizations in the high density region of strongly asymmetric nuclear matter. However, it is necessary to note, that the spin--dependent part of the Skyrme interaction at densities of relevance to neutron stars still remains to be constrained. Probably, these constraints will be obtained from the data on the time decay of magnetic field of isolated neutron stars~\\cite{PP}. In spite of this shortcoming, SLy4 and SLy5 effective forces hold one of the most competing Skyrme parametrizations at present time for description of isospin asymmetric nuclear matter at high density (together with SkM$^*$ and SGII forces for description of symmetric nuclear matter) while a Fermi liquid approach with Skyrme effective forces provides a consistent and transparent framework for studying spin instabilities in a nucleon system. In summary, we have considered the possibility of phase transitions into spin ordered states of symmetric and strongly asymmetric nuclear matter within the Fermi liquid formalism, where NN interaction is described by the Skyrme effective forces (SkM$^*$, SGII and SLy4, SLy5 potentials for the regions of vanishing and strong isospin asymmetry, respectively). In contrast to the previous considerations, where the possibility of formation of FM spin polarized states was studied on the base of calculation of magnetic susceptibility, we obtain the self--consistent equations for the FM and AFM spin order parameters and solve them in the case of zero temperature. It has been found that nuclear matter demonstrates different behavior at high densities with respect to spin fluctuations in isospin symmetric and strongly isospin asymmetric cases. In the model with SkM$^*$ and SGII effective forces symmetric nuclear matter undergoes a FM phase transition, when the spins of protons and neutrons are aligned along the same direction. In the model with SLy4 and SLy5 effective forces strongly asymmetric nuclear matter is subjected to a phase transition into the spin polarized state with the oppositely directed spins of neutrons and protons, while the state with the same direction of the neutron and proton spins does not appear. In the last case, an important peculiarity of the corresponding phase transition is the existence of long tails in the density profile of the neutron spin polarization parameter near the transition point. This means that even small admixture of protons to neutron matter leads to a considerable shift of the critical density of spin instability in the direction of low densities. In the model with SLy4 effective interaction this displacement is from the critical density $\\varrho\\approx3.7\\varrho_0$ for neutron matter to $\\varrho\\approx2.4\\varrho_0$ for asymmetric nuclear matter at the isospin asymmetry $\\alpha=0.95$, i.e. for $2.5\\%$ of protons only. As a result, the state with the oppositely directed spins of neutrons and protons appears, where protons become totally polarized in a very narrow density domain. As a consequence of this study, important questions appear, what is the value of the threshold asymmetry, at which the parallel spin ordering at small isospin asymmetry is changed to the antiparallel spin ordering at large isospin asymmetry, and do the obtained results survive for another type of an effective interaction, e.g., for Gogny effective force~\\cite{BGG,FVS} or monopole effective interaction~\\cite{RSM}? This research is in progress and will be reported elsewhere." }, "0403/astro-ph0403101_arXiv.txt": { "abstract": "Although atmospheric transmission spectroscopy of HD209458b with the Hubble Space Telescope has been very successful, attempts to detect its atmospheric absorption features using ground-based telescopes have so far been fruitless. Here we present a new method for probing the atmospheres of transiting exoplanets which may be more suitable for ground-based observations, making use of the Rossiter effect. During a transit, an exoplanet sequentially blocks off light from the approaching and receding parts of the rotating star, causing an artificial radial velocity wobble. The amplitude of this signal is directly proportional to the effective size of the transiting object, and the wavelength dependence of this effect can reveal atmospheric absorption features, in a similar way as with transmission spectroscopy. The advantage of this method over conventional atmospheric transmission spectroscopy is that it does not rely on accurate photometric comparisons of observations on and off transit, but instead depends on the relative velocity shifts of individual stellar absorption lines within the same on-transit spectra. We used an archival VLT/UVES data set to apply this method to HD209458. The amplitude of the Rossiter effect is shown to be $1.7^{+1.1}_{-1.2}$ m/sec higher in the Sodium D lines than in the weighted average of all other absorption lines in the observed wavelength range, corresponding to an increment of 4.3$\\pm$3\\% (1.4$\\sigma$). The uncertainty in this measurement compares to a photometric accuracy of 5$\\times 10^{-4}$ for conventional atmospheric transmission spectroscopy, more than an order of magnitude higher than previous attempts using ground-based telescopes. Observations specifically designed for this method could increase the accuracy further by a factor 2$-$3. ", "introduction": "\\begin{table*} \\caption{ \\label{log} The log of the observations, with in column 1 the date, in column 2 the total the exposure time, in column 3 the typical signal to noise ratio per exposure, in columns 4 and 5 the range in airmass and seeing, and in column 6 the orbital phase of HD209458b.} \\begin{tabular}{ccccccrl} Data&Observing & Exposure & S/N & Airmass & Seeing & Phase & Comments \\\\ Set&Date & Times (sec) & & & ($''$) & & \\\\ 1&05/08/2002& 15$\\times$400 & 400 & 1.407$-$1.558& $\\sim$2& $-$0.007,$+$0.021&on transit\\\\ 2&11/08/2002& 15$\\times$400 & 535 & 2.151$-$1.394&0.9$-$1.5& $+$0.663,$+$0.693&off transit\\\\ 3&12/08/2002& 15$\\times$400 & 515 & 1.391$-$2.085&0.6$-$0.9& $-$0.012,$+$0.017&on transit\\\\ 4&15/08/2002& 15$\\times$400 & 475 & 1.388$-$1.704&0.7$-$1.7& $+$0.826,$+$0.857&off transit\\\\ 5&13/09/2002& 15$\\times$400 & 485 & 1.958$-$1.385&0.8$-$1.8& $+$0.003,$+$0.026&on transit\\\\ 6&20/09/2002& 15$\\times$400 & 515 & 1.695$-$1.384&0.7$-$1.0& $-$0.009,$+$0.020&on transit\\\\ 7&22/09/2002& 15$\\times$400 & 590 & 1.446$-$1.493&0.5$-$1.0& $+$0.569,$+$0.599&off transit\\\\ \\end{tabular} \\end{table*} Since the discovery that the extra-solar planet HD209458b transits its host star (Charbonneau et al. 2000; Henry et al. 2000; Mazeh et al. 2000), many attempts have been made to detect its atmosphere using transmission spectroscopy. During a transit, light that passes from the star through the outer parts of the planet's atmosphere has impressed on it a spectrographic signature of the atmospheric constituents, which can be observed as an extra absorption on top of the stellar spectrum (Seager \\& Sasselov 2000; Brown 2001; Hubbard et al. 2001). Observations with the Hubble Space Telescope (HST) using this technique have been a great success. First the Sodium D feature was discovered by Charbonneau et al. (2002) with a relative depth of 0.02\\%, followed by the detection of very strong absorption features at a 5$-$10\\% level in the ultra-violet from Hydrogen, Oxygen and Carbon (Vidal-Madjar et al. 2003; 2004). The latter most likely come from the `exosphere' of HD209458b, an extended cometary tail of gas evaporating from the planet caused by its migration close to the star (eg. Schneider et al. 1998). In strong contrast, attempts to detect the atmosphere of HD209458b from the ground have so far been unsuccessful. So far, only upper limits of typically 1-2\\% have been reached, for the Sodium D feature in the optical (Brown et al. 2000; Bundy \\& Marcy 2000; Moutou et al. 2001), and for He, CO, H$_2$O and CH$_4$ in the near-infrared (Brown et al. 2002; Harrington et al. 2002; Moutou et al. 2003). Ground-based observations suffer greatly from the fact that transmission spectroscopy relies on the comparison of spectra on and off transit, which are necessarily taken on different nights. This limits the accuracy due to varying weather conditions and instability of the instruments. In this paper a new method for probing the atmospheres of transiting exoplanets is presented, making use of the Rossiter effect. Due to the rotation of the host star, a transiting planet will first block off light from the approaching and then from the receding parts of the stellar surface. This results in an artificial wobble in radial velocity, an effect first observed by Rossiter (1924) for the eclipsing binary $\\beta$ Lyrae. The amplitude of this signal is directly proportional to the effective size of the transiting planet, and the wavelength dependence of this effect can reveal atmospheric absorption features, in the same way as with transmission spectroscopy. The advantage of this method is that it does not rely on accurate photometric comparison of spectra in and out of transit, but instead depends on relative velocity shifts of stellar absorption lines in the same in-transit spectra. ", "conclusions": "The analysis of the UVES archive data shows that the Rossiter effect provides a powerful tool to probe the atmospheres of transiting exoplanets. Although no significant signal has been detected for Sodium D in HD209458b, it is the first time that ground-based observations result in an accuracy of $<$0.001. Previous ground-based observations have reached accuracies of only 1$-$2\\%, corresponding to an excess in the Rossiter effect of 20$-$50 m/sec, clearly inferior to the precision achieved here. The HST Transmission spectrum of HD209458b obtained by Charbonneau et al. (2002) revealed an excess in absorption in Sodium D of 0.00023 compared to the surrounding continuum, implying a relative increase of $\\sim$1.8\\%. This indicates that for this particular absorption feature an excess in the Rossiter effect of $\\sim$0.7 m/sec can be expected, making its detection very challenging for the current instrumentation. However, it is important to realise that the excess in absorption as determined with the HST was measured over a bandwidth of 12\\AA. In contrast, line-to-line variations in the Rossiter effect will be most sensitive to planetary absorption on scales of the half-power width of the stellar lines (a fraction of an \\AA ngstrom). Most models of transmission spectra (eg. Brown 2001) indicate that on this scale the absorption signal for Sodium D could be 2$-$5 times higher (1$-$3 m/sec). The archival data sets used in this paper are not optimal for measuring the Rossiter effect. First of all, since a significant fraction of the time was devoted to calibrators, typically only 60\\% of the on-transit time was spent on target, and on several occasions the periods that the amplitude of the Rossiter effect is expected to be the strongest were missed. In addition, the current analysis is significantly hampered by the fact that no or hardly any time was spent on target directly before or after each transit. This was often dictated by the timing of the transit and visibility of HD209458 from Paranal. For the most optimal transits, HD209458 can be observed from 1 hour before to 1 hour after the overall transit with the VLT. In those cases the zero-point in velocity can be determined independently from the on-transit data, and one could also correct for any possible residual drifts in radial velocity over the transit, eg. due to residual telluric absorption line features or inaccuracies in flat fielding. Overall, this should result in an increase in precision of a factor $2-3$. It means that, if the current value is believed, a 3$-$5 $\\sigma$ detection of Sodium D should be possible over a couple of transits. \\subsection*{Wavelength variations in limb-darkening} The exact shape and amplitude of the Rossiter effect are dependent on the stellar limb darkening. So far we have not taken into account that limb darkening decreases as function of wavelength. Furthermore, the limb darkening effect also changes differently over the profile of each absorption line. Here we assess the possible influence of both effects. The wavelength dependence of the limb darkening of HD209458 has been observed by Deeg et al. (2001), with $\\epsilon$ changing from 0.62 at 5900 \\AA$ $ to 0.58 at 6180 \\AA. To estimate the influence of this effect we assumed a linear wavelength dependence of $\\epsilon(\\lambda) = 1.46-\\lambda/7000\\AA$, and calculated the change in radial velocity for each line relative to that of Sodium using the model described above. By incorporating this wavelength dependent limb-darkening, the excess of the Rossiter effect in the Sodium D lines is increased by $\\sim$0.2 m/sec. The variations in limb darkening across solar absorption lines is studied in great detail by Pierce \\& Slaughter (1982) for Sodium D, and by Balthasar (1988) for a whole range of bright absorption lines, by measuring changes in the absorption line depths across the solar disk. The strength of this effect is dependent on the mechanism of formation of the line. Most absorption lines weaken towards the limb of the Sun, meaning that the limb darkening in the centre of the lines is less than in the surrounding continuum. Assuming that this effect is similar for HD209458 as for the Sun, our modeling indicates that these variation contribute at a level of again $\\sim$0.2 cm/sec. Although this is insignificant for these data, it means that future observation aimed at measuring variations in the Rossiter effect at much higher precision should take these effects into account. \\subsection*{Future applications for this method} We believe that the new method presented here to probe atmospheres of transiting exoplanets using the Rossiter effect, has great potential. Specifically targeted observations, rather than the archival data used here, can increase the accuracy further by a factor 2$-$3. Furthermore, in the near-infrared, where the atmospheric absorption features due to H$_2$O, CO, and CH$_4$ cover many stellar lines, a great improvement over current observations can be reached. Although confusion with water-vapor and methane in the earth atmosphere will complicate the analysis. In addition, anticipating the discovery of exoplanets transiting bright stars of either later spectral type (with many more stellar lines to average), or with faster stellar rotations (resulting in a Rossiter effect with a higher amplitude), a further significant increase in accuracy can be expected." }, "0403/astro-ph0403337_arXiv.txt": { "abstract": "The structure and evolution of low-mass W UMa type contact binaries are discussed by employing Eggleton's stellar evolution code \\citep{egg71,egg72,egg73}. Assuming that these systems completely satisfy Roche geometry, for contact binaries with every kind of mass ratios (0.02$\\sim$1.0), we calculate the relative radii ($R_{1,2}/A$, where $R_{1,2}$ are the radii of both stars, and $A$ the orbital separation) of both components of contact binaries in different contact depth between inner and outer Roche lobes. We obtain a radius grid of contact binaries, and can ensure the surfaces of two components lying on an equipotential surface by interpolation using this radius grid when we follow the evolution of the contact binaries. Serious uncertainties concern mainly the transfer of energy in these systems, i.e., it is unclear that how and where the energy is transferred. We assume that the energy transfer takes place in the different regions of the common envelope to investigate the effects of the region of energy transfer on the structure and evolution of contact binaries. We find that the region of energy transfer has significant influence on the structure and evolution of contact binaries, and conclude that the energy transfer may occur in the outermost layers of the common convective envelope for W-type systems, and this transfer takes place in the deeper layers of the common envelope for A-type systems. Meanwhile, if we assume that the energy transfer takes place in the outermost layers for our model with low total mass, and find that our model steadily evolves towards a system with a smaller mass ratio and a deeper envelope, suggesting that some A-type W UMa systems with low total mass could be considered as the later evolutionary stages of W-subtype systems, and that the surface temperature of the secondary excesses that of the primary during the time when the primary expands rapidly, or the secondary contracts rapidly, suggesting that W-subtype systems may be caused by expansion of the primary, or by the contraction of the secondary. ", "introduction": "It is probable that more than 50 per cent of all stars are in binary or multiple systems. An unknown, but possibly large, percentage of these systems are sufficiently close that sometime during their lifetime interacting as a result of Roche lobe overflow (RLOF). The W UMa-type contact binaries are the most common ones, comprising some 95 per cent of eclipsing variables in the solar neighborhood \\citep{shap48} or one stars in every 1000--2000 in the same spectral range \\citep{egg67}. Allowing for selection effects, W UMa stars may even contribute 1 per cent of all F and G dwarfs \\citep{vant75a}. A more recent discussion of contact binaries in the solar neighborhood is carried out by \\citet{ruc02}. He considers the complete sample of 32 EW, EB, and ellipsoidal (ELL) variables with $V<7.5$, and gives the frequency as 1 per 500 stars with $-0.5 300 {\\rm K}$) between the two components occurs in a part of the time of a cycle (lasting about 30$\\sim$35 percent time of a cycle). Almost all of the previous investigators thought that this requires there to be as many short-period binary with EB light curves as with EW light curves, and that the models for contact binaries encounter a difficulty, so called light curve paradox. \\citet{ruc02} gives 13 EWs and 5 EBs (and 14 ELLs, which have too small an amplitude to be classified as EWs or EBs). It is reasonable to identify the EWs as contact binaries and the EBs as semi-detached. The ratio of 5/13 is not much out of line with TRO theory, so we are not sure there is in fact any light curve paradox. However, the W UMa systems indeed undergo angular momentum loss without doubt, and the most likely angular momentum loss mechanism is magnetic braking \\citep{hua66,mes68}. The {\\it Einstein} X-ray observations and {\\it IUE} ultraviolet observations \\citep{eat83} showed that W UMa systems are strong sources, suggesting surface activity of the kind we observe on the Sun, and so the presence of the magnetic fields. The stellar wind would cause magnetic braking, and we will discuss the evolution of W UMa systems included the angular momentum loss in our future work. Meanwhile, we do not consider the energy source at secondary's atmosphere provided by the accreting matter from the primary at our present work. This energy source can hasten the expansion of the secondary, and shorten the time spent in the semi-detached evolution. We refer to a forthcoming work included this extra energy source for the secondary." }, "0403/astro-ph0403517_arXiv.txt": { "abstract": "{ We observed the pre--stellar core L1521F in dust emission at 1.2mm and in two transitions each of \\NTHP, \\NTDP, \\CEIO \\ and \\CSEO \\ in order to increase the sample of well studied centrally concentrated and chemically evolved starless cores, likely on the verge of star formation, and to determine the initial conditions for low--mass star formation in the Taurus Molecular Cloud. The dust observation allows us to infer the density structure of the core and together with measurements of CO isotopomers gives us the CO depletion. \\NTHP \\ and \\NTDP \\ lines are good tracers of the dust continuum and thus they give kinematic information on the core nucleus. We derived in this object a molecular hydrogen number density n(\\HH)$\\sim 10^6$\\percc \\ and a CO depletion factor, integrated along the line of sight, $f_{\\rm D}$ $\\equiv$ 9.5$\\times$10$^{-5}$/$x_{\\rm obs}$(CO) $\\sim 15$ in the central 20\\arcsec , similar to the pre--stellar core L1544. However, the $N(\\NTDP )/N(\\NTHP )$ column density ratio is $\\sim$ 0.1, a factor of about 2 lower than that found in L1544. The observed relation between the deuterium fractionation and the integrated CO depletion factor across the core can be reproduced by chemical models if \\NTHP \\ is slightly (factor of $\\sim$2 in fractional abundance) depleted in the central 3000 AU. The \\NTHP \\ and \\NTDP \\ linewidths in the core center are $\\sim$ 0.3 \\kms , significantly larger than in other more quiescent Taurus starless cores but similar to those observed in the center of L1544. The kinematical behaviour of L1521F is more complex than seen in L1544, and a model of contraction due to ambipolar diffusion is only marginally consistent with the present data. Other velocity fields, perhaps produced by accretion of the surrounding material onto the core and/or unresolved substructure, are present. Both chemical and kinematical analyses suggest that L1521F is less evolved than L1544, but, in analogy with L1544, it is approaching the ``critical'' state. ", "introduction": "The pre--stellar core L1544 has recently been the subject of much study because while apparently in hydrostatic equilibrium, there are indications that it is close to the ``critical'' state at which it will become gravitationally unstable and from which it will dynamically collapse (see discussions by Tafalla et al.~\\cite{taf98}; Ciolek \\& Basu~\\cite{cb2000}; Caselli et al.~\\cite{cas02a}; Caselli et al.~\\cite{cas02b}). If correct, this is of fundamental importance because it defines the initial conditions for the formation of a protostar, which affects many theoretical studies of low mass star formation.\\\\ Clearly trying to extrapolate general trends from a single object is difficult and a larger number of L1544-like cores (preferably with the same external environment) should be studied. Unfortunately there are rather few other objects with similar properties due to the short timescale of this phase. According to the Ciolek \\& Basu (\\cite{cb2000}) model, for example, (contraction of a disk driven by ambipolar diffusion) L1544--like properties fit the model structure of a core at times~3-$8\\times 10^4$ years prior to the collapse, after an evolution of $2.6\\times10^6$ years. Thus in that particular model, L1544 finds itself in the last few percent of its evolution prior to becoming a protostar. While this may be a somewhat too literal interpretation of the model results, it shows that ``L1544--type cores'' should be relatively rare. Further progress requires the definition of what is a L1544-like core. One answer is to use current estimates of dust emission and absorption selecting cores of dust extinction upwards of 50 mag. Another approach is to say that cores which show signs of infalling gas (as does L1544, see Williams et al.~\\cite{Wil99}; Tafalla et al.\\cite{taf98}) are ``L1544-twins''. This latter indicator is complicated by the fact that at the high densities found in the nuclei of cores similar to L1544, many molecular species and in particular CO and CS freeze--out onto dust grain surfaces (see Kramer et al.~\\cite{kra99}; Caselli et al.~\\cite{cwt99}; Bacmann et al.~\\cite{bac02}; Bergin et al.~\\cite{berg02}; J\\o rgensen et al.~\\cite{jsv02}; Tafalla et al.~\\cite{taf02}); observing such tracers implies observing the low density surrounding envelope. However, recent studies indicate that species whose abundance is linked to that of molecular nitrogen such as N$_{2}$H$^{+}$ and NH$_{3}$ (as well as their deuterated counterparts) do not condense out in the same fashion and hence can be used as tracers of the dense gas (Bergin \\& Langer~\\cite{bl97}). The extent to which this is true is debatable but it is a useful hypothesis and substantiated by the general similarity of the spatial distributions seen for example in dust emission and in maps of \\NTHP (Tafalla et al.~\\cite{taf02}, ~\\cite{tmc03}).\\\\ Caselli et al. (\\cite{cas02a},\\cite{cas02b}) have used \\NTHP \\ and \\NTDP \\ to derive the physical, chemical and kinematical properties of L1544. They found that L1544 has a central \\NTHP ~column density of $1.5 \\times 10^{13}$ cm$^{-2}$ and a column density ratio $N(\\NTDP )/N(\\NTHP )$ of 0.24. The \\NTHP \\ linewidths towards the nucleus (the dust emission peak) are roughly 0.3 km s$^{-1}$ and decrease as one goes to positions away from the center. The line of sight velocity measured in \\NTHP (1-0) and \\NTDP (2-1) shows a gradient along the minor axis of the elliptical structure seen in 1.3mm dust emission but no clear gradient along the major axis. In this paper, we will study another core in the Taurus complex, L1521F (at an assumed distance of 140 parsec) using the same approach as in our study of L1544. Repeating the L1544 study carried out by Caselli et al. (\\cite{cas02a}, \\cite{cas02b}) is important because it allows us to check to what extent L1544 is an exceptional case. In order to do this we need another source which has the same general characteristics as L1544. The source selection was made using some preliminary results we obtained in a survey carried out at the IRAM-30m telescope. L1521F stood out as being the only core in Taurus, besides L1544, with strong \\NTDP (2-1) emission compared to \\NTHP (1-0). This suggests enhanced deuterium fractionation implying an advanced evolutionary state (Caselli et al.~\\cite{cas02b}). Previous observations of this object have been carried out by Mizuno et al. (\\cite{moh94}), Onishi et al. (\\cite{omk96}), Codella et al. (\\cite{cod97}), and Lee et al. (\\cite{lom99}). Onishi et al. (\\cite{omf99}) also studied L1521F (which they call MC 27), and found a high central density, suggesting that this is the most evolved starless condensation in Taurus. L1521F was also noted by Lee et al.~(\\cite{lmt99}) as a strong infall candidate, in their survey of CS and \\NTHP \\ lines in starless cores, although later mapping of the two tracers has shown extended ``red'' asymmetry in the CS(2--1) profiles (Lee et al.~\\cite{lmt01}). In section 2 of this paper, we describe our observational procedure. In section 3 we present the observational results deriving the physical characteristics of the source and analysing its chemical and kinematical properties. In section 4 we discuss the observational results and the summary can be found in Section 5. ", "conclusions": "We have analysed physical and chemical properties of L1521F, a starless core in the Taurus Molecular Cloud, with characteristics similar to the pre--stellar core L1544. The main similarities and differences between the two cores are listed below: 1. the dust emission distributions are similar, implying a fairly closely matched density structure, with central densities of $n_0$ $\\sim$ 10$^6$ \\percc , the radius of the ``flat'' region $r_0$ = 20\\arcsec , and similar asymptotic power index $\\alpha$ (see Sect.~\\ref{sdensity} and Tafalla et al.~\\cite{taf02}). In particular, the aspect ratio is quite similar: 1.6 and 1.8 in L1521F and L1544, respectively. 2. The line width decreases with distance from the cloud center ($\\sim$ 0.3 \\kms ) to 80\\arcsec \\ away ($\\sim$ 0.25 \\kms ; see Fig.~\\ref{dv_b}), in analogy with L1544, and consistent with the predictions of ambipolar diffusion models, although any gravity--driven contraction in 2-3D is expected to get localized line broadening. The particular model which best fits the data will be investigated in the future. 3. The amount of CO freeze--out (integrated CO depletion factor $f_{\\rm D}$ = 15) is also comparable to L1544, as is the column density of \\NTHP \\ toward the dust peak ($\\simeq$ 1.5$\\times$10$^{12}$ cm$^{-2}$). 4. The deuterium fractionation toward the L1521F dust peak ($R_{\\rm deut}$ $\\sim$ 0.1) is however a factor $\\sim$ 2 smaller than in L1544, due to the (factor of 2) smaller column density of \\NTDP \\ toward L1521F. This can be understood if L1521F is less chemically evolved than L1544, with a smaller ($r < 2000$ AU) central molecular hole. 5. Unlike in L1544, the velocity field in L1521F maintains a complex structure even at the small scales traced by \\NTDP \\ and \\NTHP (3--2) (see Figs.~\\ref{fhgrad},\\ref{fdgrad}). This may be due to the presence of clumps in the central few thousand AU (as deduced by the interferometric observations of Shinnaga et al.~\\cite{sol03}), but could also be caused by normal mode pulsations, as in the case of B68 studied by Lada et al. (\\cite{lba03}). The ambipolar diffusion model with infall of Ciolek \\& Basu (\\cite{cb2000}) has problems in reproducing the whole velocity field observed across the core. This may be due to the fact that part of the observed bulk motions result from residual core contraction, as suggested by Tafalla et al.~(\\cite{tmc03}) in their study of L1517B and L1498, thus preventing a clear determination of the velocity field within the core nucleus. 6. The line profiles in L1521F show asymmetric structure, although the two peaks are not well separated as in L1544. This is consistent with smaller (factor of ~1.5) infall velocities in the central few thousand AU of L1521F. In any case, the large central density ($\\sim$ 10$^6$ \\percc ) and the evidence of central infall (broader line widths toward the center and central infall asymmetry in \\NTHP (1--0)) indicate that L1521F is another starless core on the verge of star formation, or a pre--stellar core. Although a study of a more complete sample is needed, assuming that L1544 and L1521F are the only two cores in Taurus close to dynamical collapse, and given that the total number of starless cores in Taurus is 44 (Onishi et al. \\cite{omk02}), we can argue that this process of contraction towards the ``critical'' stage, or the ``L1544--phase'', may last about five percent of the core lifetime. A more detailed analysis of \\NTHP \\ line profiles will be presented in a future paper, where the observed chemical abundances will be introduced in a non spherically symmetric Monte Carlo radiative transfer code. This is needed to understand the nature of the \\NTHP (1--0) line asymmetry, which may be caused by self--absorption plus infall, or to infall plus a molecular hole, or to the relative motion of different clumps, or to a mixture of the above possibilities." }, "0403/astro-ph0403451_arXiv.txt": { "abstract": "{The X-Ray Flash (XRF), 031203 with a host galaxy at $z=0.1055$, is, apart from GRB\\,980425, the closest $\\gamma$-Ray Burst (GRB) or XRF known to date. We have monitored its host galaxy from 1--100\\,days after the burst. In spite of the high extinction to the source and the bright host, a significant increase and subsequent decrease has been detected in the apparent brightness of the host, peaking between 10 and 33\\,days after the GRB. The only convincing explanation is a supernova (SN) associated with the XRF, SN2003lw. This is the earliest time at which a SN signal is clearly discernible in a GRB/XRF (apart from SN1998bw). SN2003lw is extremely luminous with a broad peak and can be approximately represented by the lightcurve of SN1998bw brightened by $\\sim0.55$\\,mag, implying a hypernova, as observed in most GRB-SNe. The XRF--SN association firmly links XRFs with the deaths of massive stars and further strengthens their connection with GRBs. The fact that SNe are also associated with XRFs implies that \\emph{Swift} may detect a significant population of intermediate redshift SNe very soon after the SN explosions, a sample ideally suited for detailed studies of early SN physics. ", "introduction": "} It is now firmly established that at least some long-duration $\\gamma$-ray bursts (GRBs) are accompanied by the contemporaneous explosion of a supernova \\citep[SN, e.g.][]{2003Natur.423..847H,2003ApJ...591L..17S,2003A&A...406L..33D,1998Natur.395..670G}, consistent with expectations for some models of GRBs involving the collapse of massive stars \\citep{2001ApJ...550..410M,1999ApJ...524..262M}. The lack of large numbers of SN--GRB associations may be explained, at least in part, by the difficulty of obtaining the optical spectra of SNe with redshifts usually greater than unity, against the combined backgrounds of the fading afterglow and the host galaxy. X-Ray Flashes (XRFs), a class of very soft bursts, was discovered with \\emph{BeppoSAX} \\citep{2001grba.conf...16H}. They are intense, short-lived flashes of soft X-rays of extragalactic origin \\citep{2003ApJ...599..957B,2003astro.ph.11050S,2004astro.ph..2085P}, and may be defined by a larger X-ray than $\\gamma$-ray fluence in the burst \\citep[$S_{\\rm X}/S_\\gamma>1$,][]{2003astro.ph.12634L}. The similarity in the durations of XRFs and GRBs \\citep{2001grba.conf...16H,2003A&A...400.1021B}, the continuum of spectral properties observed between the two classes \\citep{2003astro.ph.12634L}, their cosmological origins in each case \\citep{2003ApJ...599..957B,2003astro.ph.11050S}, and the similarity of their optical and X-ray afterglows \\citep{2004astro.ph..2240F,2004astro.ph..1225W}, makes it seem probable that XRFs and GRBs have a similar origin. While GRBs and XRFs are located at cosmological distances, few have been located at redshifts $<0.3$. They are: GRB\\,030329 at $z=0.1685$ \\citep[associated with SN2003dh,][]{2003Natur.423..847H,2003ApJ...591L..17S}, XRF\\,020903 with its probable host galaxy at $z=0.251$ \\citep{2003astro.ph.11050S}, GRB\\,980425 probably associated with SN1998bw at $z=0.0085$ \\citep{1998Natur.395..670G} and recently, the XRF referred to as GRB\\,031203 at $z=0.1055$ \\citep{2004astro.ph..1225W,2004astro.ph..2085P}. Of these, XRF\\,020903 had a very low peak spectral energy and a low luminosity \\citep{2004ApJ...602..875S,2003astro.ph.11050S} and GRB\\,980425 had an extraordinarily low luminosity \\citep{1998Natur.395..663K}. That GRB\\,031203 was in fact an XRF was discovered because of the detection of a transient, outwardly moving ring of X-ray emission surrounding the afterglow \\citep{2004ApJ...603L...5V}. This was interpreted as reflection of the original burst event and early afterglow off dust sheets in the Galaxy, from which strong lower limits on the prompt soft X-ray fluence were obtained \\citep{2004astro.ph..1225W}. The high extinction toward GRB\\,031203 \\citep[$(\\,\\rm E(B-V)\\approx1$,][]{1998ApJ...500..525S,2004astro.ph..2085P}, though instrumental in allowing the detection of the dust reflection halo, also hampered attempts to follow the afterglow at optical wavelengths and no optical or infrared afterglow was detected. The location of the burst is therefore determined from X-ray and radio detections. Both of these locate GRB\\,031203 unambiguously on a sub-luminous, blue, and strongly star-forming galaxy at fairly low redshift \\citep[$z=0.1055$,][]{2004astro.ph..2085P}, where the probability of a chance association with such a galaxy is not very significant \\citep{2004astro.ph..1225W,2004astro.ph..2085P}. Strong evidence for the association of XRFs with the deaths of massive stars was present in the lightcurve of XRF~030723 \\citep{2004astro.ph..2240F}, but the analysis of that burst was complicated by the lack of a redshift. Because of the low redshift there was considerable interest in attempting to discover a SN associated with GRB\\,031203, in particular since it has been found to be an XRF and the host galaxy was therefore monitored independently by a number of groups in order to find the photometric variability that would indicate a SN \\citep[e.g. this paper;][]{2003GCN..2544....1B,2003GCN..2545....1T,2004astro.ph..3510C,2004astro.ph..3608G}. This would be the first SN associated with an XRF with a spectroscopic redshift, firmly establishing the association of XRFs with the deaths of massive stars and confirming the suspected link between XRFs and GRBs. In Sect.~\\ref{observations} we describe I-band imaging observations of the host galaxy of GRB\\,031203 (HG\\,031203) over the first 100\\,days since the burst, and in Sects.~\\ref{results} and \\ref{discussion} the discovery of a SN \\citep[named SN2003lw, ][]{2004IAUC.8308....1T} associated with the XRF and the implications of this discovery. This paper supersedes an earlier preliminary report on some of these observations \\citep{2003GCN..2493....1H}. A cosmology where $H_0=75$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_\\Lambda = 0.7$ and $\\Omega_{\\rm m}=0.3$ is assumed throughout. ", "conclusions": "} We have monitored the host galaxy of the XRF, GRB\\,031203 in the near-infrared, from 1--100\\,days following the burst. In spite of the bright host galaxy and high extinction, we have discovered positive evidence of a SN, peaking $\\sim20$\\,days after the XRF and can clearly trace the early SN rise. At $z=0.1055$, this is the closest GRB/XRF-associated SN discovered so far, after SN1998bw, and the first SN associated with an XRF with known redshift. This confirms the strong case for an association between XRFs and SNe found in XRF\\,030723 \\citep{2004astro.ph..2240F}. The SN appears to have a somewhat higher peak luminosity than observed in SN1998bw, but the lightcurve is otherwise fairly similar, implying that the SN accompanying GRB\\,031203 was a hypernova. It is likely that \\emph{Swift} will detect a significant population of faint bursts like GRB\\,031203 and hence allow the study of (Type Ib/c) core-collapse SN at much earlier times than what has been possible so far; this may have a substantial impact on SN research." }, "0403/astro-ph0403498_arXiv.txt": { "abstract": "It is well known that the energy input from massive stars dominates the thermal and mechanical heating of typical regions in the interstellar medium of galaxies. These effects are amplified tremendously in the immediate environment of young massive star clusters, which may contain thousands of O and B stars. We present models of star cluster formation that attempt to account for the interplay between feedback and self-gravity in forming clusters, with application to systems ranging from the Orion Nebula Cluster, the Galactic Center clusters like the Arches, and super star clusters. ", "introduction": "Star formation in disk galaxies appears to be dominated by a clustered mode (Lada \\& Lada 2003), occurring in regions of the disk that are gravitationally unstable (Martin \\& Kennicutt 2001). The most unstable mass scales are about $1.3\\times 10^7 (\\sigma_{\\rm gas}/6\\;{\\rm km\\:s^{-1}})^4 (\\Sigma_{\\rm gas}/10\\;M_\\odot {\\rm pc}^{-2})^{-1} M_\\odot$, roughly consistent with the observed cutoff of the Galactic Giant Molecular Cloud (GMC) mass function (Williams \\& McKee 1997). GMCs are gravitationally bound (Solomon et al. 1987), and there is a comparable amount of atomic gas associated with these structures (Blitz 1990). Altogether, a significant fraction ($\\sim 1/2$) of the total gas mass inside the solar circle is organized into bound structures. The lifetimes of GMCs are not well known. Blitz \\& Shu (1980) used observational and theoretical arguments to infer a GMC lifetime of a few times $10^7$~yr; their argument based on photoevaporation of the clouds was confirmed in a more extensive calculation by Williams \\& McKee (1997). The rocket effect associated with the photoevaporation displaces the molecular gas from regions of active star formation at the same time the gas is being transformed from molecular form to atomic and ionized form. Leisawitz, Bash \\& Thaddeus (1989) found that open star clusters older than about $\\sim 10$~Myr were not associated with molecular clouds, which is consistent either with post-star-formation cloud lifetimes shorter than this age or with relative velocities of star clusters and their parent clouds of about $10\\:{\\rm km\\:s^{-1}}$. Elmegreen (2000) presented a number of arguments for star formation to occur in about $1-2$ dynamical timescales, particularly from the small age spreads in clusters such as the Orion Nebula Cluster and from the statistics of the presence of young stars in regions of dense gas. Hartmann (2003) has argued for such rapid star formation in Taurus. On the other hand, the observation that the angular momentum vectors of GMCs in M33 are small and are both pro- and retro-grade (Rosolowsky et al. 2003) may indicate a relatively long lifetime so that angular momentum can be shed by magnetic braking (e.g., Mestel 1985) and/or cloud-cloud collisions in a shearing disk (Tan 2000). Most star formation in GMCs is concentrated in {\\it clumps} that occupy a relatively small fraction of the volume and have a small fraction of the mass. Within clumps are over-dense regions that we refer to as {\\it cores} that form individual stars or binaries (Williams, Blitz, \\& McKee 2000). Star-forming clumps are identified by $\\rm H_2O$ masers, outflows, and far infrared and radio continuum emission (Plume et al. 1997; Hunter et al. 2000; Sridharan et al. 2002; Beuther et al. 2002; Zhang et al. 2002; Mueller et al. 2002; Shirley et al. 2003). Most of the 350~$\\rm \\mu m$ emission maps of Shirley et al. (2003) have morphologies consistent with quasi-spherical, virialized distributions. Typical properties of clumps are masses $M\\sim 100-10^4\\: M_\\odot$, diameters $\\sim 1$~pc and surface densities $\\Sigma\\sim 1\\:{\\rm g\\:cm^{-2}}$. Clump properties are very similar to those of the smaller infrared dark clouds (IRDCs) (Carey et al. 1998), which are dense, cold regions embedded in GMCs. Thus, understanding the origin of IRDCs is crucial for progress in the fields of both star cluster formation and galaxy formation and evolution. Carey et al. noted nongaussian emission line profiles of $\\rm H_2CO$ from 9 out of 10 IRDCs. This may indicate short formation timescales via a triggering mechanism. IRDC morphologies are also more filamentary than the sub-mm emission maps of star-forming clumps, which suggest that the latter represent a more evolved stage. Possible triggers for IRDC formation include compression of parts of GMCs (which are likely to contain clumpy substructure) by shock waves driven by cloud-cloud collisions (Scoville, Sanders \\& Clemens 1986; Tan 2000), spiral density waves (Sleath \\& Alexander 1996), ionization fronts (Elmegreen \\& Lada 1977; Thompson et al. 2004), supernovae (e.g. Palous, Tenorio-Tagle \\& Franco 1994) or stellar winds (e.g. Whitworth \\& Francis 2002). Clumps that were previously pressure-confined and gravitationally stable may suddenly become unstable. The contraction of the cloud can be halted by the onset of star formation (McKee 1989). We expect that the protocluster will come into an approximate equilibrium in which pressure support balances self-gravity. Some pieces of evidence in support of this view are the approximately spherical morphologies of star-forming clumps (Shirley et al. 2003); the timescales of star cluster formation estimated from outflow momentum generation rates (Tan \\& McKee 2002); and the empirical age spreads of stars in young clusters, such as the Orion Nebula Cluster (Palla \\& Stahler 1999), which are quite long compared to the dynamical or free-fall timescales of clumps, $\\bar{t}_{\\rm ff}=(3\\pi/32G\\bar{\\rho})^{1/2} =1.0\\times10^5 (M/4000M_\\odot)^{1/4}\\Sigma^{-3/4}\\:{\\rm yr}$. This last point requires comment, since Elmegreen (2000) uses the Orion Nebula Cluster as an example of star formation on a dynamical time. He estimated the density in the cluster prior to star formation as $n_{\\rm H}=1.2\\times 10^5$ cm$^{-3}$, corresponding to a free-fall time of $1.25\\times 10^5$ yr. If the star formation occurred over a time $t_{\\rm sf}=10^6$ yr, then $\\eta\\equiv t_{\\rm sf}/\\bar t_{\\rm ff} \\simeq 8\\gg 1$. He used the dynamical time $t_{\\rm dyn}\\equiv R/\\sigma$ to compare with $t_{\\rm sf}$. If the virial parameter $\\alpha_{\\rm vir}\\equiv 5\\sigma^2 R/GM\\sim 1$, as observed in star-forming regions (McKee \\& Tan 2003), then $R/\\sigma =2.0\\bar t_{\\rm ff}/\\alpha_{\\rm vir}^{1/2} \\simeq 2.5\\times 10^5$ yr. For $t_{\\rm sf}\\simeq 10^6$~yr, star formation in the Orion Nebula Cluster was rapid, but not so rapid that a quasi-equilibrium treatment is invalid. ", "conclusions": "" }, "0403/astro-ph0403384_arXiv.txt": { "abstract": "We present a direct detection of the growth of large-scale structure, using weak gravitational lensing and photometric redshift data from the COMBO-17 survey. We use deep $R$-band imaging of two $0.5\\times0.5$ square degree fields, affording shear estimates for over 52000 galaxies; we combine these with photometric redshift estimates from our 17 band survey, in order to obtain a 3-D shear field. We find theoretical models for evolving matter power spectra and correlation functions, and fit the corresponding shear correlation functions to the data as a function of redshift. We detect the evolution of the power at the $7.7\\sigma$ level given minimal priors, and measure the rate of evolution for $0 10^{-10} \\ecs$; Aharonian et al. 2003) over the synchrotron component ($f_s \\simeq 2 \\times 10^{-11} \\ecs$; Costamante et al. 2001). It also must be taken into account that the bright gamma-ray state of this object is likely to be accompanied by increased synchrotron emission, which would decrease the $L_c/L_s$ ratio to values previously observed in HBL. With respect to the blazar sequence \\cite{fossati}, RX J1211+2242 is an HBL with remarkably strong gamma-ray emission, making it an interesting object for further studies at energies above $\\sim 1 \\rm \\, MeV$. The interesting region where the inverse Compton branch becomes dominant over the synchrotron branch lies in the energy range covered by the {\\it INTEGRAL} instruments ($20 - 8000 \\keV$). Given the X-ray spectrum shown by RX J1211+2242 in the {\\it BeppoSAX} data, it will not be easily detectable by this mission, even when assuming that the power law with $\\alpha_X \\simeq 1.0$ extends up to several hundred keV. We have performed simulations that show that a 500 ksec observation would give a $< 5 \\sigma$ detection by the SPI spectrometer. However, the upcoming gamma-ray missions {\\it GLAST} \\cite{GLAST} and {\\it AGILE} \\cite{AGILE}, with their improved sensitivity, should be able to detect RX J1211+2242 in the EGRET band. Even though this source is likely to be a strong TeV emitter, it will not be seen by most Cherenkov telescopes as the high energy emission of an object at redshift $z = 0.455$ will be suppressed by interaction of the TeV photons with the extragalactic infrared background \\cite{kneiske}. Nevertheless the {\\it MAGIC} telescope \\cite{MAGIC} with its low energy threshold of $\\sim 30 \\, \\rm GeV$ is suitable to observe the peak of the IC component making RX J1211+2242 one of the most promising targets in the gamma-ray sky." }, "0403/astro-ph0403572_arXiv.txt": { "abstract": "We derive absolute dimensions of the early B-type detached eclipsing binary V453\\,Cygni (B0.4\\,IV + B0.7\\,IV, $P=3.89$\\,d), a member of the open cluster NGC\\,6871. From the analysis of new, high-resolution, spectroscopy and the $UBV$ light curves of Cohen (1974) we find the masses to be $14.36 \\pm 0.20$\\Msun\\ and $11.11 \\pm 0.13$\\Msun, the radii to be $8.55 \\pm 0.06$\\Rsun\\ and $5.49 \\pm 0.06$\\Rsun, and the effective temperatures to be $26\\,600 \\pm 500$\\,K and $25\\,500 \\pm 800$\\,K for the primary and secondary stars, respectively. The surface gravity values of $\\logg = 3.731 \\pm 0.012$ and $4.005 \\pm 0.015$ indicate that V453\\,Cyg is reaching the end of its main sequence lifetime. We have determined the apsidal motion period of the system to be $66.4 \\pm 1.8$\\,yr using the technique of Lacy (1992) extended to include spectroscopic data as well as times of minimum light, giving a density concentration coefficient of $\\log k_2 = -2.226 \\pm 0.024$. Contaminating (third) light has been detected for the first time in the light curve of V453\\,Cyg; previous analyses without this effect systematically underestimate the ratio of the radii of the two stars. The absolute dimensions of the system have been compared to the stellar evolution models of the Granada, Geneva, Padova and Cambridge groups. All model sets fit the data on V453\\,Cyg for solar helium and metal abundances and an age of $10.0 \\pm 0.2$\\,Myr. The Granada models also agree fully with the observed $\\log k_2$ once general relativistic effects have been accounted for. The Cambridge models with convective core overshooting fit V453\\,Cyg better than those without. Given this success of the theoretical predictions, we briefly discuss which eclipsing binaries should be studied in order to further challenge the models. ", "introduction": "\\label{introduction} Theoretical models of high-mass stars are very difficult to construct, due to a number of poorly understood astrophysical phenomena which become important at higher stellar masses. Convective core overshooting is an essential ingredient of stellar models and can have a large effect on the lifetimes and luminosities of high-mass stars. The degree of overshooting may depend on metallicity (Cordier \\etal\\ 2002, Palmieri \\etal\\ 2002) and stellar mass (Young \\etal\\ 2001). Stellar rotation is known to cause enhanced mixing in the interiors of stars which can mimic a small amount of convective overshooting and strongly affect the evolution of stars (Maeder \\& Meynet 2000). Mass loss is also an important effect (Maeder 1997) but is usually represented in theoretical models by a parameterisation such as that of Reimers (1975) (Woo \\etal\\ 2003). Until recently there was a discrepancy between masses derived from spectroscopic observations and inferred from evolutionary models (Herrero \\etal\\ 1992). Recent advances in theoretical modelling, and the increased sophistication of observational techniques, appears to have removed this mass discrepancy (Hilditch, 2004), but confirmation of this requires detailed comparisons between models and as many physical properties of individual stars as possible. Detached eclipsing binaries (dEBs) are a vital source of observational data on the radii and masses of high-mass stars (Andersen 1991) but studies of such stars are hindered by the rarity of useful systems, the difficulty of measuring accurate velocities from spectra of rapidly-rotating multiple stars (Sana, Rauw \\& Gosset 2001), and the large amount of telescope time needed to obtain light curves which are complete over the whole orbital period. This has caused a shortage of data on eclipsing binaries with masses above 10\\Msun\\ and sufficiently large orbital periods to be detached and therefore to have evolved as single stars. Whilst dEBs can provide accurate observed stellar masses, radii, effective temperatures and equatorial rotational velocities, the ages, chemical compositions and distances of such systems are not, in general, directly observable. This means that when fitting observed quantities to theoretical predictions, there are several free parameters which can be adjusted to find the best fit to the observational data. This makes it very difficult to assess the success of important but more subtle physical effects such as rotational mixing, mass loss, convective overshooting and even the type of turbulent convective theory. Eclipsing binaries in open clusters and associations can be analysed to derive accurate masses, radii, temperatures and rotational velocities. Their membership of a star cluster can also provide knowledge of the age, chemical composition, and distance of the dEB, allowing a more complete physical description of a single stellar system. This reduces the number of free parameters adjustable when fitting theoretical models to observations, which allows a discriminate test of the representation of different physical effects and indeed the success of different sets of stellar evolutionary models. For example, from the analysis of two main sequence dEBs which are members of the open cluster h\\,Persei (NGC\\,869), Southworth, Maxted \\& Smalley (2004, hereafter Paper\\,I) were able to provide a first estimate of the bulk metallicity of one of the most important and well-studied open clusters in the Northern Hemisphere. \\subsection{V453\\,Cygni in NGC\\,6871} \\label{v453cyg} \\begin{table} \\begin{center} \\caption{\\label{tablephotdata} Identifications, location, and combined photometric parameters for the V453\\,Cygni eclipsing system. \\newline {\\bf References:} (1) Cannon \\& Pickering (1923); (2) Argelander (1903); (3) H{\\o}g \\etal\\ (1998); (4) Hoag \\etal (1961); (5) Popper (1980); (6) Zakirov (1992); (7) Cohen (1969); (8) Reimann (1989).} \\begin{tabular}{lr@{}lr} \\hline \\hline & & V453 Cygni & Reference \\\\ \\hline Henry Draper number & & HD 227696 & 1 \\\\ Bonner Durchmusterung & & BD\\,+35\\degr 3964 & 2 \\\\ Hoag number & & NGC 6871 13 & 3 \\\\ \\hline $\\alpha_{2000}$ & & 20 06 34.967 & 4 \\\\ $\\delta_{2000}$ & + & 35 44 26.28 & 4 \\\\ Spectral type & & B\\,0.4\\,IV + B\\,0.7\\,IV & 5 \\\\ \\hline $V$ & & 8.285 & 6 \\\\ $B-V$ & + & 0.179 & 6 \\\\ $U-B$ & $-$ & 0.61 & 6 \\\\ $V-R$ & + & 0.254 & 6 \\\\ $\\beta$ & & 2.590 & 7,8 \\\\ \\hline \\hline \\end{tabular} \\end{center} \\end{table} \\begin{table*} \\begin{center} \\caption{\\label{pubspecorbits} Published spectroscopic orbits of V453\\,Cygni. BMM97 originally fitted their data with a circular orbit. We have refitted their radial velocities with an eccentric orbit to increase the accuracy of our determination of the apsidal motion (see section~\\ref{perioddet} for details). A colon after a number indicates that it is uncertain. Quantities without quoted errors or a colon were not determined by that investigation. When quantities are given separately for each star we have quoted a weighted mean of the two values. Symbols have their usual meanings. Times are written as (HJD $-$ 2\\,400\\,000). \\newline $^*$\\,The reference time, $T_0$, refers to a time of periastron passage, not a time of minimum light.} \\begin{tabular}{l r@{\\,$\\pm$\\,}l r@{\\,$\\pm$\\,}l r@{\\,$\\pm$\\,}l r@{\\,$\\pm$\\,}l r@{\\,$\\pm$\\,}l r@{\\,$\\pm$\\,}l } \\hline \\hline & \\mc{Pearce} & \\mc{Abt, Levy and} & \\mc{Popper and} &\\mc{Simon and} & \\mc{BMM97} & \\mc{BMM97} \\\\ & \\mc{(1941)} & \\mc{Gandet (1972)} & \\mc{Hill (1991)}&\\mc{Sturm (1994)}& \\mc{} & \\mc{(our solution)} \\\\ \\hline $P$ (days) & \\mc{3.87972} & \\mc{3.8890} & \\mc{3.8898128} & \\mc{3.88982309} &\\mc{3.8898128} &\\mc{3.889825} \\\\ $T_0$ (HJD) &30231.0843&0.0543$^*$ & \\mc{40495.027$^*$} & \\mc{39340.099} & \\mc{36811.7296} &48141.82&0.01$^*$& 48500.64 & 0.66$^*$ \\\\ $K_{\\rm A}$ (\\kms) & 181.8 & 1.13 & \\mc{152:} & 171 & 1.5 & 171.7 & 2.9 & 173.2 & 1.3 & 173.7 & 1.4 \\\\ $K_{\\rm B}$ (\\kms) & 237.4 & 2.78 & \\mc{} & 222 & 2.5 & 223.1 & 2.9 & 213.6 & 3.0 & 212.4 & 3.4 \\\\ $e$ & 0.07 & 0.007 & \\mc{0.05:} &\\mc{0.0} &\\mc{0.0} & \\mc{0.0} & 0.011 & 0.015 \\\\ $\\omega$ (degrees) & 175.2 & 5.06 & \\mc{99:} & \\mc{} & \\mc{} & \\mc{} & 88.6 & 6.0 \\\\ \\Vsys\\ (\\kms) & $-$15.0&0.94 & \\mc{$-$22.7:} &\\mc{$-$14} & \\mc{$-$7:} & $-17.6$ & 1.0 & $-$18.0 & 1.6 \\\\ \\hline \\hline \\end{tabular} \\end{center} \\end{table*} V453\\,Cygni is a high-mass dEB with an orbital period of 3.89\\,days. Its membership of the young open cluster NGC\\,6871 means that its age and distance can be found independently. The primary component of V453\\,Cyg is approaching the terminal age main sequence (TAMS) and its large radius causes the eclipses to be total, allowing a very accurate determination of the radii of both stars. Table~\\ref{tablephotdata} contains identifications and some photometric properties of the system. The eclipsing nature of V453\\,Cyg was discovered by Wachmann (1939) and an early spectroscopic orbit was calculated by Pearce (1941). A period study by Cohen (1971) provided a determination of the orbital longitude of periastron, $\\omega$, inconsistent with that derived by Pearce. In a period study by Wachmann (1973) this was correctly interpreted as apsidal motion. Wachmann derived an apsidal period of $U = 72$ years using measurements of the time differences between several groups of adjacent primary and secondary eclipses. A more recent period study, including parabolic and periodic terms, was undertaken by Rafert (1982). Excellent photoelectric $UBV$ light curves were observed by Wachmann (1974) and analysed using the Russell-Merrill method (Russell \\& Merrill 1952) which involves the procedure of rectification. His work contains a plot of the light curves adjusted for the effects of orbital eccentricity and apsidal motion using parameters updated from that of his previous work, but the data themselves have so far been unobtainable. It is possible that they are in an unlabelled file in the IAU Variable Star Archives (Breger 1988), but no record of them exists at Hamburg Observatory, where the light curves were observed (A.\\ Reiners, private communication). Cohen (1974) published complete photoelectric $UBV$ light curves which contain fewer datapoints and more observational scatter than those of Wachmann (1974). He analysed these using the Russell-Merrill method but stated that his observations were not definitive. They have since been analysed by Cester \\etal\\ (1978) using the light curve analysis code {\\sc wink} (Wood 1971). This is the only previous photometric study to use modern techniques. A recent investigation using photoelectric $UBVRI$ light curves has been published by Zakirov (1992). He analysed his light curves using the ``direct machine method of Lavrov (1993)'', which is based on rectification. The results of the four photometric analyses of V453\\,Cyg are substantially in agreement about the basic photometric parameters of the system. Recent spectroscopic orbits have been published by Popper \\& Hill (1991), Simon \\& Sturm (1994) and Burkholder, Massey \\& Morrell (1997, hereafter BMM97). These results are collected in Table~\\ref{pubspecorbits}. Simon \\& Sturm used seven spectra to demonstrate their spectral disentangling algorithm, which decomposes observed composite spectra into the separate spectra of two stars. The total secondary eclipse of V453\\,Cyg allowed them to directly compare their disentangled primary spectrum with a spectrum observed during secondary eclipse. Disentangling can be used to determine accurate orbital semiamplitudes (Hynes \\& Maxted 1998, Harries \\etal\\ 2003) but it is not clear if there is a robust method by which to estimate uncertainties in the derived quantities. BMM97 derived a good spectroscopic orbit from 25 high signal-to-noise spectra and compared the dEB to models to investigate the discrepancy at higher masses between models and observations. The rotational velocities of the components of V453\\,Cyg were determined by Olson (1984) to be $107 \\pm 9$\\kms\\ and $97 \\pm 20$\\kms\\ for the primary and secondary stars respectively. A preliminary single-lined spectroscopic orbit was also given by Abt, Levy \\& Gandet (1972). An abundance analysis of V453\\,Cyg was undertaken by Daflon \\etal\\ (2001) using both LTE and non-LTE calculations. The results suggest that V453\\,Cyg has a slightly sub-solar metallicity. These authors derived an effective temperature of 29\\,100\\,K using the $Q$ parameter based on $UBV$ magnitudes (Johnson 1957), and a surface gravity of $\\logg = 4.45$ (\\cms) from profile fitting of the H$\\gamma$ 4340\\,\\AA\\ spectral line. Both values are larger than expected and inconsistent with previous analyses. Their surface gravity value is in fact somewhat larger than appropriate for the ZAMS, and is inconsistent with significant main sequence evolution. \\subsection{NGC\\,6871} \\label{ngc6871} The open cluster NGC\\,6871 is a concentration of bright OB stars which forms the nucleus of the Cyg\\,OB3 association (Garmany \\& Stencel 1992). This makes it an important object for the study of the evolution of high-mass stars. The cluster itself has been studied photometrically several times but its sparse nature means determination of its physical parameters is difficult. $UBV$ photometry of the 30 brightest stars was published by Hoag \\etal\\ (1961). Crawford, Barnes \\& Warren (1974) observed 11 stars in the Str\\\"omgren $uvby$ system and 24 stars in the Crawford $\\beta$ system, finding significantly variable reddening and a distance modulus of 11.5\\,mag. This $uvby\\beta$ photometry was extended to 40 stars by Reimann (1989), who found reddening $\\Eby$ with a mean value of 0.348\\,mag and an intracluster variation of about 0.1\\,mag. His derived distance modulus of $11.94 \\pm 0.08$\\,mag and age of 12\\,Myr are both greater than previous literature values. Massey, Johnson \\& DeGioia-Eastwood (1995) conducted extensive $UBV$ CCD photometry of 1955 stars in the area of Cyg\\,OB3. Their values of distance modulus, $11.65 \\pm 0.07$, and of reddening, $\\EBV = 0.46 \\pm 0.03$\\,mag with individual values between 0.04 and 1.11 mag, agree well with previous determinations. They find an age of 2 to 5 Myr for stars with spectral types earlier than B0 but give evidence for a significant spread of stellar ages in the cluster. Whilst the highest-mass unevolved cluster members have main sequence lifetimes of 4 to 5 Myr, NGC\\,6871 contains evolved 15\\Msun\\ stars despite their main sequence lifetimes being of the order of 11\\,Myr. ", "conclusions": "\\label{discussion} We have derived the absolute dimensions of the components of the high-mass detached eclipsing binary V453\\,Cygni, a member of the open cluster NGC\\,6871. Effective temperatures were found using the helium ionisation balance derived from high-resolution spectra, which also suggest an enhanced photospheric helium abundance relative to solar. Radial velocities were derived from the spectra using only the weak spectral lines and the {\\sc todcor} two-dimensional cross-correlation algorithm. The apsidal motion rate of the system has been determined using an extended version of the photometric method of Lacy (1992), which includes times of minimum light and spectroscopic determinations of eccentricity and $\\omega$. The apsidal period is well constrained, and allow the derivation of eccentricity and $\\omega$ to a greater accuracy than possible with the light curves and radial velocity curves. We have reanalysed the $UBV$ light curves of Cohen (1974) in order to determine the radii of the components of the dEB. The best-fitting parameters include a small amount of third light, which was previously undetected. Robust parameter uncertainties were derived using bootstrapping, allowing us to quantify and illustrate the effect of correlations between different photometric parameters. The ratio of the radii and the amount of third light are strongly correlated, due to their dependence on the depths of the eclipses; previous photometric studies which did not include third light are systematically biased towards values of the stellar radii which are 1\\% higher and 5\\% lower for primary and secondary respectively. The accurate absolute dimensions presented here allow V453\\,Cyg to be added to the list of dEBs with the best-determined values of mass, radius and effective temperature (Andersen 1991). However, our analysis would clearly be much improved with better observational data. The inclusion of only a few new times of minima would greatly increase the accuracy of the results of the apsidal motion analysis, and more accurate rotational velocities would allow a more accurate derivation of the internal structure constant $\\log k_2$. A definitive spectroscopic orbit will require observations with a higher signal-to-noise than those presented here, and should give masses determined to accuracies of better than 1\\%. Definitive light curves of the system would allow determination of the limb darkening coefficients for both stars, providing an important test of model atmosphere codes. The absolute masses, radii and effective temperatures of the components of V453\\,Cyg have been compared to several stellar models in the mass--radius and $\\log\\Teff$--\\logg\\ planes, assuming the same age for both stars. Not only is there impressive agreement betwen different theoretical models, all model sets are able to fit the observational data for a solar helium and metal abundance. Moreover, the Granada models (Claret 1995) provide a perfect match to the observed apsidal motion rate once the relativistic contribution has been subtracted from the overall effect. Stellar models have for a long time appeared to predict that the central condensations of stars are lower than that found using observations of apsidal motion (e.g., Clausen, Gim\\'enez \\& Scarfe 1986, Barembaum \\& Etzel 1995). This apparent discrepancy has been reduced by the discovery that the internal structure constants change significantly through a star's evolution. The current generation of theoretical models, incorporating OPAL opacity data (Rogers \\& Iglesias 1992), are in good agreement with observations. It is noticeable that some observers have not removed the general relativistic effect from their observed $\\log k_2$ values before comparison with theory; in many cases this will have a negligible effect but for the stars of V453\\,Cyg it causes about 6\\% of the observed apsidal motion, changing $\\log k_2$ by an amount similar to its uncertainty. The normal helium abundance implied by stellar model fits also conflicts with the slight overabundance noted in our spectral synthesis analysis. We note that the photospheric helium abundance is not directly comparable to the initial internal helium abundance used in model calculations. Fits to the Cambridge stellar models support the inclusion of a moderate amount of overshooting in most stellar evolutionary models. Whilst models without overshooting were able to fit the masses and radii of the stars, the predicted effective temperatures are slightly lower than that determined from the helium ionisation balance. The stellar models were extremely successful in fitting the absolute dimensions and effective temperatures of a high-mass slightly-evolved dEB, with component masses and radii differing by ten and twenty-five times their combined uncertainties, respectively. For observational stellar astrophysicists, this fact implies that we must either observe systems so thoroughly that their masses and radii are known to accuracies of 0.5\\% and the effective temperatures to 2\\%, or target particular types of stars to critique the success of one set of stellar models compared to another. Such targets include low-mass, high-mass, pulsating, and Population II stars, as well as eclipsing systems found in Local Group galaxies. Eclipsing binaries in open clusters can satisfy this requirement if the cluster they belong to is well-studied or otherwise interesting (Paper\\,I), and further observations are being undertaken towards this goal." }, "0403/astro-ph0403434_arXiv.txt": { "abstract": "{In the proceedings of this, and of several recent close binary conferences, there have been several contributions describing smoothed particle hydrodynamics simulations of accretion disks. It is apposite therefore to review the numerical scheme itself with emphasis on its advantages for disk modelling, and the methods used for modelling viscous processes. } \\addkeyword{accretion, accretion disks} \\addkeyword{binaries:close} \\addkeyword{line:profiles} \\addkeyword{x-rays:binaries} \\addkeyword{hydrodynamics} \\addkeyword{methods:numerical} \\begin{document} ", "introduction": " ", "conclusions": "One of the great advantages of the smoothed particle hydrodynamics technique is its simplicity. An SPH code is rarely more than a thousand lines long. This, in combination with established methods for incorporating new physics into the algorithm, makes it an ideal tool for studying systems where the important physics is uncertain. In the case of accretion disk research, we can not yet describe with any confidence the form of the anomalously large dissipation. We can however make use of SPH simulations to investigate various postulated forms for disk viscosity (e.g. the viscoelastic description developed by Ogilvie, 2003). The encouraging message is thus that if there is viscosity in an SPH simulation of a disk, it is there specifically to model the unknown process driving accretion in close binaries. It is without doubt possible to have shocks and a prescribed viscous dissipation term in the same simulation. We conclude by referring the reader to presentations of SPH disk simulations at this conference by Hayasaki, Kunze, Manson, Truss and Okazaki." }, "0403/astro-ph0403328_arXiv.txt": { "abstract": "A point mass at the center of an ellipsoidal homogeneous fluid is used as a simple model to study the effect of rotation on the shape and external gravitational field of planets and stars. Maclaurin's analytical result for a homogenous body is generalized to this model. The absence of a third order term in the Taylor expansion of the Maclaurin function leads to further simple but very accurate analytical results connecting the three observables: oblateness ($\\epsilon$), gravitational quadrupole ($J_{2}$), and angular velocity parameter ($q$). These are compared to observational data for the planets. The moments of inertia of the planets are calculated and compared to published values. The oblateness of the Sun is estimated. Oscillations near equilibrium are studied within the model. \\\\ \\\\ ", "introduction": "The rotation induced oblateness of astronomical bodies is a classical problem in Newtonian and celestial mechanics (for the early history, see Todhunter \\cite{todhunter}). It has twice played an important role in the history of science. In the early eighteenth century measurements indicated a prolate shape of the Earth, in strong conflict with the Newtonian prediction. This was later shown to be wrong by more careful measurements by Maupertuis, Clairaut, and Celsius in northern Sweden in 1736. Then, in 1967, measurements of the solar oblateness were published and according to these it was much larger than the Sun's surface angular velocity would explain. The confirmation of general relativity by Mercury's perihelion precession would then be lost. Also this problem is now gone and the modern consensus is that the solar oblateness is too small to affect this classic test of general relativity \\cite{will,godier,laskar}. The subject of the flattening of rotating astronomical bodies is thus quite mature. The classical theory is due mainly to Clairaut, Laplace, and Lyapunov. Also Radau, Darwin, de Sitter and many others have made important contributions. More recent accounts of the theory can be found in, for example, Jeffreys \\cite{jeffreys}, Zharkov et al.\\ \\cite{zharkov}, Cook \\cite{cook}, Moritz \\cite{moritz} and, partly, in Chandrasekhar \\cite{chandrasekhar}. Some pedagogical efforts can be found in Murray and Dermott \\cite{murray}, or in Kaula \\cite{kaula}. As is plain from these references the theory is quite involved. Only the unrealistic assumption that the body is homogeneous gives compact analytical results. Otherwise a specified radial density distribution is needed and one resorts to cumbersome series (multipole) expansions, or purely numerical methods, for quantitative results. Here we will present analytical results based on the assumption that the body consists of a central point mass surrounded by a homogeneous fluid, the so called point core model. By varying the relative mass of the fluid and the central point particle one can interpolate between the extreme limits of Newton's homogeneous body and the Roche model \\cite{kippenhahn} with a dominating small heavy center. The point core model goes back to the work of G.~H.\\ Darwin. More recently it has been used to study the shape of outer planet moons, see Hubbard and Anderson \\cite{hubbard}, Dermott and Thomas \\cite{dermott}. Apart from the basic point core approximation (i) several further approximations are assumed here. These are: (ii) that the shape is determined by hydrostatic equilibrium and (iii) that the shape is ellipsoidal. These are not consistent. According to Hamy's theorem, see Moritz \\cite{moritz}, the exact shape is not ellipsoidal, so we regard an ellipsoidal shape as a constraint and find the equilibrium shape, among these, by minimizing the energy. A further approximation (iv) is the neglect of differential rotation. On the other hand fixed volume (``bulk incompressibility'') need not be assumed; the equilibrium volume problem separates from the shape problem. In spite of these approximations, which are standard in the literature, the mathematics can be quite involved. In this article I hope to clarify and simplify it as much as possible. Mathematically our model then becomes a three degree of freedom mechanical system for which we can calculate the kinetic and potential (gravitational plus centrifugal) energies exactly. Multipole expansions in terms of spherical harmonics are not needed. The three degrees of freedom correspond to the three semi-axes of the ellipsoid ($a, b, c$) but these are transformed to three generalized coordinates that describe size (or volume), $R$, spheroidal, $\\xi$, and triaxial, $\\tau$, shape changes, see Eq.\\ (\\ref{eq.def.xi.eta}). The statics problem of equilibrium shape is solved by minimizing the potential energy, $U(R,\\xi,\\tau)$ for a fixed $R$, Eq.\\ (\\ref{eq.energy.xi.tau}). For slow rotation the shape will not be triaxial so $\\tau=0$. Finding the shape, or flattening, is then only a matter of minimizing a dimensionless potential energy, \\begin{equation} \\label{eq.dim.less.pot.en.xi} u(\\xi)=\\psi(\\xi)-\\frac{1}{2}k \\xi^2 , \\end{equation} given by the last two terms of Eq.\\ (\\ref{eq.energy.xi}). Here $\\psi(\\xi)$ is defined by Eq.\\ (\\ref{eq.basic.phi.xi}) and plotted in Fig.\\ \\ref{FIG1} while $k$ is a constant that depends on rotational parameter $q$ and dimensionless moment of inertia $\\kappa^2$, Eq.\\ (\\ref{eq.kappa2.nu}). The root $\\xi(k)$ of the equation $\\psi'(\\xi) -k\\xi =0$ thus gives the flattening. This root is to very high accuracy given by \\begin{equation} \\label{eq.xi.root.of.k} \\xi(k)=\\left[1-(7/2)\\left(\\sqrt{1+15 k/7} - 1\\right) \\right]^{-1/6} , \\end{equation} see Eqs.\\ (\\ref{eq.maclaurin.approx.inv}) and (\\ref{eq.eccentricity.relation.xi}), for the rotational parameters that can be found in the solar system. From this result, simple formulas, Eqs.\\ (\\ref{eq.maclaurin.approx.relation.q}) -- (\\ref{eq.maclaurin.approx.relation.epsilon}), relating the observables, rotation parameter, $q$, gravitational quad\\-ru\\-pole, $J_{2}$, and excentricity squared, $e^2$, are obtained. These appear to be, partly, new, and their usefulness is demonstrated by comparing with empirical data for the Sun and the rotating planets of the solar system. Variational methods have been used before to study similar problems, see for example Abad et al.\\ \\cite{abad}. Denis et al.\\ \\cite{denis1} have pointed out that variational methods generally fail to provide estimates of their accuracy. Therefore the agreement of our formulas with empirical data, as demonstrated in Table /cite(table), is important and demonstrates that our model catches the essential physics of rotational flattening. Finally small amplitude oscillations near the equilibrium are investigated, starting from the Lagrangian \\begin{equation} \\label{eq.lagrangeian} L=T(R,\\xi,\\tau,\\dot R, \\dot\\xi,\\dot\\tau)-U(R,\\xi,\\tau), \\end{equation} for our three degree of freedom model system. This gives useful insight into the physics of free stellar or planetary oscillations and their coupling to rotation. Our approximations are, however, too severe for these results to be of quantitative interest. ", "conclusions": "Some apparently new results relating to the classic theory of the figure of rotating bodies have been presented. The basic model, a point mass at the center of a homogeneous fluid, is characterized by their mass ratio, and interpolates between the limits of an ellipsoidal homogeneous fluid and a body dominated by a small central mass concentration. It allows simple analytic treatment but is still flexible enough to correctly describe the essential hydrostatics of real rotating planets as well as stars. Such models are always useful, especially when one wishes to analyze, compare, and understand large numbers of observational data. In spite of the simplicity there is no perturbation order to which the results are valid; the nonlinearity of the basic equations can be retained. To be more precise Eq.\\ (\\ref{eq.maclaurin.approx}) shows that the formulas are valid to seventh order in the eccentricity (within the basic model with its simplified mass distribution). As demonstrated by the numerical experiments on Jupiter and Saturn data, this is essential. In fact Table~\\ref{table} indicates that the geometric oblateness of the surface equipotential surface of Mars and Uranus determined from observed $q$ and $J_{2}$ values using (\\ref{eq.maclaurin.approx.relation.epsilon}) probably are more reliable than current observational data. Finally the dynamics of the model reveals the essentials of the coupling and rotational splitting of the most basic free oscillations of a planet without recourse to expansion in spherical harmonics.\\\\ \\\\ \\noindent {\\bf Acknowledgement} Constructive comments from Dr.\\ John D.\\ Anderson on a previous version of this manuscript are gratefully acknowledged." }, "0403/astro-ph0403487_arXiv.txt": { "abstract": "The recent detection of variable infrared emission from Sagittarius A*, combined with its previously observed flare activity in X-rays, provides compelling evidence that at least a portion of this object's emission is produced by nonthermal electrons. We show here that acceleration of electrons by plasma wave turbulence in hot gases near the black hole's event horizon can account both for Sagittarius A*'s mm and shorter wavelengths emission in the quiescent state, and for the infrared and X-ray flares, induced either via an enhancement of the mass accretion rate onto the black hole or by a reorganization of the magnetic field coupled to the accretion gas. The acceleration model proposed here produces distinct flare spectra that may be compared with future coordinated multi-wavelength observations. We further suggest that the diffusion of high energy electrons away from the acceleration site toward larger radii might be able to account for the observed characteristics of Sagittarius A*'s emission at cm and longer wavelengths. ", "introduction": "The observation of stellar motions within light-days of Sagittarius A*, a compact radio source at the Galactic Center \\citep{Balick74}, has provided compelling evidence that this source is the radiative manifestation of a $\\sim$ four million solar mass black hole \\citep{Schodel02, Ghez031}. The recently detected infrared emission and flare activity have provided an additional evidence that this source is powered by a hot gas accreting onto the black hole \\citep{Baganoff01, Goldwurm03, Baganoff032, Porquet03, Zhao04, Ghez03}. The quasi-periodic near-infrared variability may be an indication that the gas flared up before spiraling into the black hole \\citep{Genzel03}. It is now generally agreed that the radio and infrared emission, and the flares, are likely produced by nonthermal high energy electrons (Liu \\& Melia 2001; Genzel et al. 2003; see also Mahadevan 1998). However, the exact nature of the mechanism responsible for the acceleration of the electrons has not been addressed. This has given rise to diverse interpretations of the observations with assumed spectra of the accelerated electrons \\citep{Markoff01, Liu02a, Nayakshin03, Yuan03}. In this letter, we show that the mechanism producing high-energy particles in solar flares works equally well in hot plasmas near the black hole. The solar flare model is based on a second order Fermi acceleration process or a stochastic acceleration (SA) of particles by interacting resonantly with plasma waves or turbulence (PWT) generated via an MHD dissipation process (see e.g. Miller \\& Ramaty 1987; Hamilton \\& Petrosian 1992; Petrosian \\& Liu 2004). In Sagittarius A*, nonthermal particles can be produced by the turbulence expected to be induced by the magneto-rotational instability in the accretion torus (Balbus \\& Hawley 1991; Melia, Liu \\& Coker 2001). The radiation emitted at the acceleration site can explain the quiescent state mm and shorter wavelength observations. Solar flares are energized by the process of magnetic reconnection during the dynamical evolution of the coronal magnetic field. Similar processes in Sagittarius A* can release energy in a small region and produce what we call a local event. A global flare can be induced by an MHD fluctuation in the accretion torus or an enhancement of the accretion rate onto the black hole. The emission spectra from these two energization mechanisms are quite different, and can explain the distinct flare behaviors. In \\S\\ \\ref{SA}, we outline the theory of SA. Its application to Sagittarius A* is presented in \\S\\ \\ref{sgra}. The main results of this letter are summarized in \\S\\ \\ref{dis}, where we also discuss consequences of energetic electrons escaping the acceleration site. If such electrons can diffuse toward larger radii, they may account for Sagittarius A*'s emission at cm and longer wavelengths and has the potential to explain the observed linear and circular polarization characteristics, and variabilities of this source. ", "conclusions": "\\label{dis} Based on the theory of stochastic particle acceleration, we have built a model to account for Sagittarius A*'s emission at mm and shorter wavelengths. The quiescent state emission is attributed to electrons accelerated by turbulence in a magnetized accretion torus and flares can be produced via two distinct mechanisms. The IR and X-ray flares with harder spectrum are likely induced by a local MHD process, such as magnetic reconnection. Global fluctuations generally produce flares with IR and X-ray spectral indexes close to their corresponding value in the quiescent state. The model not only accounts for the varied spectra of the flares, but also explains their relative occurrence rate observed at IR and X-ray. Radio emission at longer wavelengths cannot be produced within such a small emission region \\citep{Liu01a}. This is not surprising, given that the radio emission is variable on a longer time scale of tens of hours to one week, suggesting that this radiation is produced at larger radii than what we have been considering here \\citep{Zhao04, Zhao03, Zhao93, Bower02}. Very interestingly, our model suggests a significant outflow of high-energy electrons \\citep{Liu02b}, which may very well be the particles that eventually produce Sagittarius A*'s cm spectrum via synchrotron emission on a spatially larger scale. In the quiescent state (Model A), the power carried away by electrons with $\\gamma>100$ (electrons with lower energy may be trapped by the gravitational potential of the black hole and will not diffuse to larger radii) is about $2\\times10^{37}$ ergs s$^{-1}$, which is more than enough to power the observed radio emission, whose luminosity is about $10^{34-35}$ ergs s$^{-1}$. (Note that in this model, the mass accretion rate is $\\sim 10^{18-19}$g s$^{-1}$, for which a few percent of the dissipated gravitational energy is carried away by the outwardly diffusing electrons; see the caption of Table \\ref{tab:mods}.) For the {\\it XMM-Newton} flare on October 3 2002 (Model C), the total energy carried away by the escaping high-energy electrons is about $7\\times 10^{40}$ ergs which for a few days can sustain the radio flare observed 13 hours after the X-ray event \\citep{Zhao04}. This picture also explains the weak correlation between the {\\it Chandra} X-ray flares and the radio emission from Sagittarius A* \\citep{Baganoff032}, because the flux of high energy electrons produced during these flares is much smaller than that for the {\\it XMM-Newton} flare; no strong enhancement of radio emission is expected following a {\\it Chandra}-type of X-ray flare. One of the intriguing properties of Sagittarius A*'s radio emission is that it is circularly polarized below $100$ GHz, even though no linear polarization has been observed there \\citep{Bower02}. However, in the mm and sub-mm range only strong linear polarization is observed \\citep{Aitken00, Bower03}. The mm/sub-mm polarization is likely associated with the structure of the magnetic field at small radii \\citep{Agol00, Melia00, Bromley01}, which is consistent with the spectral formation at these wavelengths discussed in this paper. The observed circular polarization may be due to anisotropy of the escaping electrons as they are transported along magnetic field lines toward larger radii under the influence of synchrotron losses \\citep{McTiernan90}. Relativistic electrons beamed along magnetic field lines produces synchrotron emission with significant circular polarization while the degree of linear polarization can be suppressed by irregularities in the source magnetic field \\citep{Epstein73a, Epstein73b}. Work to demonstrate this feature self-consistently is in progress, and the results will be reported elsewhere. The model we have presented here shows promise in being able to account not only for the spectral characteristics of Sagittarius A*, but also for its polarization characteristics and time variation properties. Finally, we emphasize that although we suggested that $\\eta$, $k_{\\rm max}/\\Omega_{\\rm e}$, $\\alpha$, and $a$ may not change significantly over time, and have fixed them in all the models discussed above, given the dramatic changes of the physical conditions during a flare state, these parameters may also vary slightly, giving rise to different emission spectra than those presented here. Indeed, if the infrared upper limits reported by Hornstein et al. (2002) are confirmed, one then needs to decrease the high frequency cutoff of the synchrotron emission by adjusting these parameters. This will be investigated in a more comprehensive study of particle acceleration in Sagittarius A*." }, "0403/astro-ph0403164_arXiv.txt": { "abstract": "We re-formulate cosmological perturbations in the decaying cold dark matter model, and calculate cosmological microwave background (CMB) anisotropies. By comparing our theoretical predictions with recent observational data from the Wilkinson Microwave Anisotropy Probe (WMAP), we derive a new bound on the abundance and lifetime of decaying dark matter particles. We show that the data of WMAP alone do not prefer the decaying cold dark matter model: the lifetime is constrained to $\\Gamma^{-1} \\ge 123$ Gyr at $68\\%$ C.L. ($52$ Gyr at 95.4$\\%$ C.L.) when cold dark matter consists only of such decaying particles. We also consider a more general case in which cold dark matter consists of both stable and decaying particles, and show that the constraint generalizes down to $\\Omega_{\\rm DDM}h^2 \\la -0.5(\\Gamma^{-1}/1{\\rm Gyr})^{-1}+0.12$ for $\\Gamma^{-1} \\ge 5\\mbox{Gyr}$ at $95.4\\%$ C.L.. These bounds are robust and widely applicable, because they are derived only from gravitational effects and therefore do not rely on the details of decay channels or decay products. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403352_arXiv.txt": { "abstract": "We compute the properties of dark matter halos with mass $10^{6.5}-10^9\\msun$ at redshift $z=6-11$ in the standard cold dark matter cosmological model, utilizing a very high resolution N-body simulation. We find that dark matter halos in these mass and redshift ranges are significantly biased over matter with a bias factor in the range $2-6$. The dark matter halo mass function displays a slope of $2.05\\pm 0.15$ at the small mass end. We do not find a universal dark matter density profile. Instead, we find a significant dependence of the central density profile of dark matter halos on halo mass and epoch with $\\alpha_0=0.4-1.0$; the high-mass ($M\\ge 10^8\\msun$) low-redshift ($z\\sim 6$) halos occupy the high end of the range and low-mass ($M\\sim 10^{7}\\msun$) high-redshift ($z\\sim 11$) halos occupy the low end. Additionally, for fixed mass and epoch there is a significant dispersion in $\\alpha_0$ due to the stochastic assembly of halos. Our results fit a relationship of the form $\\alpha_0=0.75((1+z)/7.0)^{-1.25}(M/10^7\\msun)^{0.11(1+z)/7.0}$ with a dispersion about this fit of $\\pm 0.5$ and no systematic dependence of variance correlated with environment. The median spin parameter of dark matter halos is $0.03-0.04$ but with a large lognormal dispersion of $\\sim 0.4$. Various quantities are tabulated or fitted with empirical formulae. ", "introduction": "The reionization epoch is now within the direct observational reach thanks to rapid recent observational advances in two fronts --- optical quasar absorption from Sloan Digital Sky Survey (SDSS) (Fan \\etal 2001; Becker \\etal 2001) and the Wilkinson Microwave Anisotropy Probe (WMAP) experiment (Kogut \\etal 2003). The picture painted by the combined observations, perhaps not too surprisingly, strongly suggests a complex cosmological reionization process, consistent with the double reionization scenario (Cen 2003). It may be that this is the beginning of a paradigm shift in our focus on the high redshift universe: the star formation history of the early universe can now be observationally constrained. It thus becomes urgent to theoretically explore galaxy and star formation process at high redshift in the dark age ($z\\ge 6$). In the context of the standard cold dark matter model it is expected that stars within halos of mass $10^7-10^9\\msun$ at high redshift play an important, if not dominant, role in determining how and when the universe was reionized. Furthermore, these fossil halos may be seen in the local universe as satellites of giant galaxies. This linkage may potentially provide a great leverage to nail down the properties of the high redshift galaxies. In this paper, as a step towards understanding galaxy formation at high redshift, we investigate the properties of dark matter halos at $z\\ge 6$, using very high resolution TPM N-body (Bode \\etal 2001; Bode \\& Ostriker 2003) simulations. While there is an extensive literature on properties of halos at low redshift, there is virtually no systematic study of dark halos at $z\\ge 6$. The LCDM simulation has a comoving box size of $4h^{-1}$Mpc with $512^3=10^{8.2}$ particles, a particle mass of $m_p=3.6\\times 10^4 \\ h^{-1}\\msun$, and comoving gravitational softening length of $0.14 \\ h^{-1}$kpc. These resolutions allow us to accurately characterize the properties of halos down to a mass $10^{6.5} \\ h^{-1}\\msun$ (having about $100$ particles within the virial radius). The outline of this paper is as follows. The simulation details are given in \\S 2. In \\S 3 we quantify properties of dark matter halos in the mass range $10^{6.5}-10^9\\msun$, including the mass function, bias and clustering properties, density profile distribution, angular momentum spin parameter distribution, internal angular momentum distribution and peculiar velocity distribution. We conclude in \\S 4. ", "conclusions": "Using a high resolution TPM N-body simulation of the standard cold dark matter cosmological model with a particle mass of $m_p=3.57\\times 10^4 \\ h^{-1}M_\\odot$ and a softening length of $\\epsilon=0.14 \\ h^{-1}$kpc in a $4h^{-1}$Mpc box, we compute various properties of dark matter halos with mass $10^{6.5}-10^9\\msun$ at redshift $z=6-11$. We find the following results. \\noindent (1) Dark matter halos at such small mass at high redshifts are already significantly biased over matter with a bias factor in the range $2-6$. \\noindent (2) The dark matter halo mass function displays a slope at the small end $2.05\\pm 0.15$. \\noindent (3) The central density profile of dark matter halos are found to be in the range $(0.4-1.0)$ well fitted by $\\alpha_0=0.75((1+z)/7.0)^{-1.25}(M/10^7\\msun)^{0.11(1+z)/7.0}$ with a dispersion of $\\pm 0.5$, in rough agreement with the theoretical arguments given in Ricotti (2003) and Subramanian \\etal (2000). \\noindent (4) The median spin parameter of the dark matter halos is $0.03-0.04$ but with a lognormal dispersion of $\\sim 0.4$. The angular momentum profile at the small end is approximately linear with the fraction of mass in a halo having specific angular momentum less than a certain value is roughly $0.5$ times the ratio of that value over the average specific angular momentum of the halo. \\noindent (5) The dark matter halos move at a typical velocity in excess of $30-40$km/s." }, "0403/astro-ph0403678_arXiv.txt": { "abstract": "When the ion mean free path much exceeds the Larmor radius in a plasma, the viscous stress tensor is altered dramatically, and depends only upon quantities measured along the field lines. This regime corresponds to typical interstellar medium conditions in galaxies and protogalaxies, even if the magnetic field is extremely weak, with a negligible Lorentz force on all scales of interest. In this work, the only role of the magnetic field is to channel angular momentum transport along its lines of force. We show that differential rotation in such a gas is highly unstable, with a maximum growth rate exceeding that of the magnetorotational instability. The regime of interest has been treated previously by plasma kinetic methods. Where there is overlap, our work appears to be in agreement with the kinetic results. The nonlinear outcome of this instability is likely to be a turbulent process, significantly augmenting the magnetorotational instability, and important to the initial phases of the amplification of small galactic magnetic fields. ", "introduction": "The magnetorotational instability, or MRI, has become central to our understanding of turbulent angular momentum transport in accretion disks (e.g. Balbus 2003). The instability is important even when (indeed, especially when) the magnetic energy density is small compared with the thermal energy density. It is this feature, the fact that the stability of the gas is hypersensitive to the presence of subthermal magnetic fields of any geometry, that endows the MRI with its special significance. In fact, the MRI is merely one manifestation of much more general behavior. By imparting new degrees of freedom to a fluid, magnetic fields allow free energy gradients to become sources of instability, with important consequences for a variety of astrophysical systems: accretion disks become turbulent when the angular velocity (not angular momentum) decreases outward, and stratified dilute gases are destabilized when the temperature (not entropy) decreases upwards \\citep{b01}. % The thermal destabilization caused by magnetic field is especially noteworthy because it is engendered by what seems at first a purely dissipative process: thermal diffusivity. It is the extreme anisotropy of the conductivity tensor parallel and perpendicular to the field lines that lies at the heart of the instability. By channeling heat only along lines of magnetic force, small-amplitude ripples along initially isothermal field lines grow into large fluid displacements parallel to the temperature gradient. This is not, as is sometimes thought, analogous to classical double-diffusive instabilities, such as ocean layer ``salt-fingering''\\footnote{See Menou, Balbus, \\& Spruit (2004) for a true salt-fingering analogy involving the MRI.}. Rather, it depends wholly upon the properties of anisotropic conductivity. In this paper we show that the anisotropy introduced into the viscous stress tensor by a weak magnetic field sharply destabilizes dilute astrophysical disks, even without Lorentz forces appearing in the fluid equations. By ``dilute,'' we mean the limit in which the ion Larmor radius is small compared to a mean free path, which in turn is small compared with the characteristic macroscopic length scales of the disk. We shall refer to the resulting viscous stress tensor in this limit as Braginskii viscosity \\citep{b65}. The Braginskii limit is appropriate for interstellar and galactic disks (especially protogalactic disks, cf. Malyshkin \\& Kulsrud 2002). A remarkable property of the instability is that in the absence of Lorentz forces, when the viscous diffusivity much exceeds the resistive diffusivity, rapid growth times are associated with arbitrarily high wave numbers. Except for isolated field geometries (e.g. precisely azimuthal), there is no formal high wavenumber dissipation of the linear magneto-viscous instability. By way of contrast, for the ultra-weak magnetic fields considered here, the wavelength of maximum growth in the classical MRI would be strongly damped by an ordinary isotropic viscosity. (Note that in real systems, at sufficiently high wave numbers, Lorentz forces will ultimately stabilize.) The magnetic stability of a dilute plasma may also be studied using a plasma kinetic approach. This offers rigor in a problem where it is clearly of some benefit, but at a price of greater mathematical complexity. \\citet{qdh02} analyzed the MRI in the collisionless regime, and noted that the character of the instability changes when the pressure tensor becomes significantly anisotropic, with growth rates in excess of the classical MRI maximum. In the current work, we have in essence abstracted the anisotropic components of the pressure, and labeled them as a viscosity tensor. More recently, Sharma, Hammett \\& Quataert (2003) reanalyzed the kinetic problem using a Krook collision operator, and showed that the collisional and collisionless behavior are on the same branch of the dispersion relation. In the short mean free path limit, the pressure tensor associated with the equations used by Sharma et al. 2003 reduces to a scalar pressure plus the dominant parallel components of the Braginskii viscosity (Snyder, Hammett, \\& Dorland 1997). The current work, a purely gasdynamical treatment of the problem, affords relative mathematical simplicity and a physically transparent interpretation. In either its kinetic or gasdynamical guise, this vigorous instability may be important in the early stages of magnetic field amplification in disk galaxies, when densities are low, temperatures are relatively high, and the field is likely to be very weak. Quataert et al.\\ (2002) and Sharma et al.\\ (2003) discuss applications to black hole accretion flows of low radiative efficiency. An outline of the paper is as follows. Section 2 is a discussion of the regime of the applicability of this work and a presentation of the magnetic viscosity formalism. Section 3 is the heart of the paper, formulating and solving the problem, and checking the validity of the results. Section 4 is a discussion of the applicability of our results to galactic magnetism, and section 5 summarizes our findings. ", "conclusions": "The instability presented in this paper is generic and powerful. Its most rapidly growing behavior is exhibited at large vertical wavenumbers, and even in the presence of finite resistivity, it is not easily quenched. Long radial and azimuthal wavelengths are highly unstable, imparting large scale coherence in the disk plane already in the linear stages of instability. In this section we discuss the possibility that this mechanism could be an important part of the process that gives rise to galactic magnetic fields. Generally, galactic magnetic amplification processes are divided into two principal categories \\citep{beck96}: (1) differential wind-up of a primordial field (Kulsrud 1986); and (2) dynamo amplification of a similar seed field. The latter category itself consists of at least two very distinct mechanisms; (2a) a classical $\\alpha\\Omega$ dynamo (Parker 1979), and (2b) small scale turbulent amplification (Schekochihin et al. 2004). The weak field limit of these mechanisms are thought to be kinematic in nature. It is difficult to see how the process outlined in this paper could be unimportant, at least quantitatively, to any of these processes. Simple wind-up of a seed field by differential rotation fails to incorporate properly the true dynamics of either the MRI or the magneto-viscous instability described here. Differential rotation and essentially any field geometry is highly unstable, resulting in large radial motions and ultimately MHD turbulence. The shearing of radial fields of course readily occurs, but it is not the primary amplification process of a differentially rotating system. Traditional $\\alpha\\Omega$ galactic dynamo models rely on the presence of favorable properties of interstellar turbulence to generate the required mean helicity. Curiously, these theories of magnetic field amplification generally do not incorporate weak-field MHD instabilities, despite the latter's obvious potential benefits (e.g. Brandenburg et al. 1995; Hawley, Gammie, \\& Balbus 1996). Conversely, without something like the MRI or the magneto-viscous instability, it is not so obvious that turbulence will generally amplify the field, at least at low to moderate values of ${\\cal P}$. An example of dissipative behavior is seen in a simulation of Hawley et al.\\ (1996). The combination of hydrodynamical shear-layer turbulence plus a magnetic field lead not to dynamo activity, but to field energy loss. This result emerged despite the fact that calculation was done not in the kinematic limit, but fully in the MHD regime. However, the combination of local Coriolis, tidal, and Lorentz forces dramatically altered the dynamo properties of the ensuing turbulence: rapid and significant field amplification was observed, driven by the MRI turbulence, in a hydrodynamically stable background. More recently, the notion that non-helical homogeneous small scale turbulence may play a key role in galactic dynamos, particularly in the early stages, has been investigated in a series of numerical simulations (Schekochihin et al. 2004; see also Zeldovich et al. 1984, Kulsrud \\& Anderson 1992), which include a scalar viscosity and study the non-kinematical regime. The idea is that certain magnetic field configurations (termed ``winning'' by Schekochihin et al.) align themselves smoothly with the stretching direction of the strain tensor of the turbulent flow, but fluctuate along the corresponding null axis, so that the work done on the field by fluid element stretching is not undone by relaxation. A sort of turbulent ratchet thereby ensues, growing the field with enormous efficiency at small scales. Numerical simulations of homogeneous white noise forcing conducted by Schekochihin et al. 2004 resulted in exponential field amplification, but, interestingly, only in the regime of large ${\\cal P}$. What the relationship is between these small scale structures and a galactic scale field has yet to be established. Further discussion of the pros and cons of this mechanism would take us too far afield, but we may note that the generation of coherent magnetic field structure on scales larger than the disk scale height requires time scales at least as long as $1/\\Omega$. And while it may well be that intrinsic interstellar turbulence plays a role in the amplification of galactic magnetic fields, we emphasize here a phenomenon that is certainly unavoidable: differential rotation. In fact, there is no reason to restrict oneself to galactic radius scales: if sub-galactic, differentially rotating turbulent vorticies are also present, magneto-viscous modes should be seen on these scales as well. The key point is that a kinematic approach will miss this process, which is intrinsically MHD. One compelling origin of the seed field for the galactic amplification process is the stars themselves (Biermann 1950, Rees 1993), in which both battery processes and a truly powerful, rapid dynamo are likely to be present. A 0.1 G azimuthal surface field diluted in a stellar wind to interstellar scales would give rise to a seed field of $\\sim 2\\times 10^{-9}$G. With $T\\simeq 10^4$ K, $n\\simeq 1$ cm${}^{-3}$, one finds $$ \\omega_{ci}\\tau_{ci} \\simeq 10, \\quad {\\cal P} \\sim 10^{11} $$ and the regime of the magneto-viscous instability is valid. The ratio of gas pressure to magnetic pressure is $3.5\\times 10^7$, an extremely weak field by this measure. The linear amplification factor per orbit of the magnetic energy, as noted in \\S 3.2, is huge: $5.2\\times 10^7$. The makings of an MHD dynamo would seem to be present. If a combination of magneto-viscous and magneto-rotational processes is to be a viable candidate for galactic field production, it needs to be shown that (1) the bulk of the field energy emerges in the largest scales; and (2) the saturated field energy density can grow to levels comparable to the thermal pressure. Definitive answers will require a numerical treatment, but in the meantime the following points may be considered. A magnetic energy spectrum dominated by the largest scales appears to be a universal outcome of MRI simulations (e.g. Hawley, Gammie, \\& Balbus 1995), despite the fact that the most unstable modes occur at scales much smaller than the scale height. In such calculations, the available dynamical range is limited. To the extent it can be measured, however, the inertial range is Kolmogorov-like in the magnetic energy. In hydrodynamics, Kolmogorov scaling for the kinetic energy power spectrum is universal, if the dissipation rate per unit volume is the sole constant characterizing the cascade, as often seems to be the case. Universal processes may also be at work in MHD turbulence (Kraichnan 1963; Goldreich \\& Sridhar 1997). An important caveat, however, is that most MHD turbulence studies have been based on wave-wave interactions (which may be of secondary importance in linearly unstable rotating systems), and they have yet to address the role of a magnetized viscosity. Indeed, a magnetized viscosity may be crucial to understanding why interstellar fields are thermal strength (or even slightly above), whereas all numerical MRI simulations to date have yielded subthermal saturated field strengths. It has been argued (Balbus \\& Hawley 1998) that the outcome may be well be sensitive to ${\\cal P}$. The point is that large scale reconnection proceeds relatively easily in simulations when the resistive scale is comparable to or larger than the viscous scale. In ``ideal MHD,'' both scales are grid based. But matters are likely to be very different if the viscous scale is, say, five orders of magnitude larger. In that case, enormous viscous stresses would occur in the course of setting up a reconnection front, and prevent its formation. With reconnection stifled, MHD turbulence could amplify the field to its natural dynamical limit: the thermal energy density. Beyond this point, buoyant effects would make it difficult for a suprathermal field to remain in the disk and be further amplified (Parker 1979). Large Prandtl number simulations have in fact recently begun (Schekochihin et al. 2004), and treatments of the anisotropic Braginskii viscosity and conductivity have yet to be attempted. With the resistivity scale hidden below the viscous dissipation scale, there is no effective small scale sink for the magnetic field, and is therefore not surprising to note that Schekochihin et al.\\ find that the magnetic energy spectra increases with wave number on subviscous scales, before it is ultimately cut-off. By way of sharp contrast, the ideal MRI simulations noted above find a monotonically decreasing energy spectrum for the magnetic field (Hawley, Gammie, \\& Balbus 1995). The magneto-viscous instability is a powerful and general mechanism to amplify weak magnetic fields in galaxies, or in hot dilute plasmas more generally, and is worthy of more detailed study. Of particular interest would be a suite of numerical simulations designed to isolate the differences between high Prandtl number turbulent dynamos relying on MHD and differential rotation, those based on random forcing only, and those containing both." }, "0403/astro-ph0403022_arXiv.txt": { "abstract": "Spatial distributions of energy deposited by an extensive air shower in the atmosphere through ionization, as obtained from the CORSIKA simulation program, are used to find the fluorescence light distribution in the optical image of the shower. The shower image derived in this way is somewhat smaller than that obtained from the NKG lateral distribution of particles in the shower. The size of the image shows a small dependence on the primary particle type. ", "introduction": "\\label{intro} The fluorescence method of extensive air shower detection is based on recording light emitted by air molecules excited by charged particles of the shower. For very high energies of the primary particle, enough fluorescence light is produced by the large number of secondaries in the cascading process so that the shower can be recorded from a distance of many kilometers by an appropriate optical detector system \\cite{greisen,Bal}. As the amount of fluorescence light is closely correlated to the particle content of a shower, it provides a calorimetric measure of the primary energy. The particles in an air shower are strongly collimated around the shower axis. Most of them are spread at distances smaller than several tens of meters from the axis, so that when viewed from a large distance, the shower resembles a luminous point on the sky. Therefore, a one-dimensional approximation of the shower as being a point source might be adequate in many cases regarding the shower reconstruction. For more detailed studies, however, the spatial spread of particles in the shower has to be taken into account. This is especially important for nearby showers, where the shower image, i.e.~the angular distribution of light recorded by a fluorescence detector (FD), may be larger than the detector resolution. The image of a shower has been studied in Ref.~\\cite{sommers}, where it was shown that for a disk-like distribution of the light emitted around the shower axis, the shower image has a circular shape, even when viewed perpendicular to the shower axis. Analytical studies including lateral particle distributions parameterized by the Nishimura-Kamata-Greisen (NKG) function or estimates based on average particle distributions taken from CORSIKA~\\cite{heck} were discussed in Ref.~\\cite{dgora} and Ref.~\\cite{giller}, respectively. In this paper, detailed Monte Carlo simulations of the shower image based on the spatial energy deposit distributions of individual showers are performed. By using the energy deposit of the shower particles as calculated by CORSIKA~\\cite{markus3}, the previous simplified assumption of a constant fluorescence yield per particle is avoided. Assuming a proportionality between the fluorescence yield and ionization density, the light emitted by each segment of the shower is determined. A concept is developed to treat the shower as a three-dimensional object, additionally taking into account the time information on photons arriving at the FD. In contrast to previous analytical studies, shower fluctuations as predicted by the shower simulation code are preserved and studied. Propagation of the light towards the detector, including light attenuation and scattering in the atmosphere is simulated, so that the photon flux at the detector is calculated. The resulting distribution of photons arriving simultaneously at the detector, i.e. the shower image, is compared to results obtained by using the NKG approximation of particle distribution in the shower. The comparison is performed for different shower energies and different primary particles. In particular, it is checked whether the shower width depends on the primary particle type. \\\\ \\\\ The plan of the paper is the following: definition of the shower width and algorithms of fluorescence light production are described in Section 2. In Section 3 the size of shower image in the NKG and CORSIKA approaches is calculated and its dependence on primary energy, zenith angle and primary particle is discussed. Conclusions are given in Section 4. ", "conclusions": "Shower image simulations more accurate than available until now are presented, which incorporate a more realistic distribution of fluorescence light emitted by the shower. The image simulations are based on distributions of energy deposited by the shower in air as derived from CORSIKA. A comparison of the size of the shower image obtained using CORSIKA and that given by the NKG function was made for different energies and primary particles. To a first approximation, the results of these two completely independent methods (analytical versus Monte Carlo) show quite reasonable agreement. The image spot size derived from CORSIKA is smaller by about 15\\% compared to the NKG approximation. This difference is mainly due to the differences in lateral particle distributions in the NKG and CORSIKA approximation. The energy deposit distribution from CORSIKA leads to a dependence of the size of shower image on the primary particle, so that studies of the shower image may be helpful for the primary particle identification. {\\it Acknowledgements.} We would like to thank R. Engel and F. Nerling for fruitful discussions and careful reading of the manuscript. This work was partially supported by the Polish Committee for Scientific Research under grants No. PBZ KBN 054/P03/2001 and 2P03B 11024 and by the International Bureau of the BMBF (Germany) under grant No. POL 99/013." }, "0403/astro-ph0403508_arXiv.txt": { "abstract": "An elementary kinematic model for emission produced by relativistic spherical colliding shells is studied. The case of a uniform blast-wave shell with jet opening angle $\\theta_j \\gg 1/\\Gamma$ is considered, where $\\Gamma$ is the Lorentz factor of the emitting shell. The shell, with comoving width $\\Delta r^\\prime$, is assumed to be illuminated for a comoving time $\\Delta t^\\prime$ and to radiate a broken power-law $\\nu L_\\nu$ spectrum peaking at comoving photon energy $\\e_{pk,0}^{\\prime}$. Synthetic GRB pulses are calculated, and the relation between energy flux and internal comoving energy density is quantified. Curvature effects dictate that the measured $\\nu F_\\nu$ flux at the measured peak photon energy $\\e_{pk}$ is proportional to $\\e^3_{pk}$ in the declining phase of a GRB pulse. Possible reasons for discrepancy with observations are discussed, including adiabatic and radiative cooling processes that extend the decay timescale, a nonuniform jet, or the formation of pulses by external shock processes. A prediction of a correlation between prompt emission properties and times of the optical afterglow beaming breaks is made for a cooling model, which can be tested with Swift. ", "introduction": "In the collapsar scenario for GRBs, pulses in GRB light curves are thought to be produced by collisions between relativistic shells ejected from a central engine (see \\citet{zm04} for a recent review). The interception of a more slowly moving shell by a second shell that is ejected at a later time, but with faster speed and larger Lorentz factor, produces a shock that dissipates internal energy to energize the particles that emit the GRB radiation. This scenario is widely considered to explain pulses in GRB light curves \\citep{kps97,dm98}. Studies of pulses are important to decide if GRB sources require engines that are long-lasting or impulsive \\citep{dm03}, with important implications for the nature of the central engine, which is often argued to be a newly formed black hole powered by the accretion of a massive dense torus. Here we construct an elementary kinematic model for colliding shells, assumed spherical and uniform within jet opening angle $\\theta_j$. This is the sort of jet that \\citet{fra01} discuss regarding the standard energy reservoir result, where jet opening angles are inferred from the time of achromatic spectral breaks in optical afterglow light curves. We also perform this study in order to quantify the curvature constraint of a spherically emitting shell traveling with bulk Lorentz factor $\\Gamma$, which implies that the shell radius \\begin{equation} r \\approx 2\\Gamma^2 c t_{var}/(1+z)\\; \\label{r} \\end{equation} in order to produce variability on timescale $t_{var}$ \\citep{rl79,fmn96}. This study also quantifies the rate at which flux decays at a given energy due to curvature effects, and the range of validity of the approximate relation \\begin{equation} \\Phi_E \\cong c r^2 u_0^\\prime \\Gamma^2/d_L^2\\; \\label{phie} \\end{equation} between internal comoving energy density $u_0^\\prime$ and observed energy flux $\\Phi_E$, where $d_L$ is the luminosity distance (See Appendix A). The accuracy of this relation is important to quantify $\\gamma\\gamma$ opacity constraints \\citep{ls01,der04} applied to GRB pulses as measured with the GRB monitor and Large Area Detector on GLAST\\footnote{glast.gsfc.nasa.gov}, as well as to make estimates of photomeson production in GRB blast waves \\citep{wb97}. If curvature effects dominate the late-time emission in GRB pulses, then a unique relation is found whereby the value of the $\\nu F_\\nu$ peak flux $f_{\\epk}$(in cgs units of ergs cm$^{-2}$ s$^{-1}$) at peak photon energy $\\e_{pk}$ decays in proportion to $\\propto\\e_{pk}^3$. This relation is generally not observed in long, smooth GRB pulses studied by \\citet{br01}, who find power-law decays $f_{\\epk}\\propto \\epk^\\zeta$, with $0.6 \\lesssim \\zeta \\lesssim 3$. Remarkably, values of $\\zeta$ for different pulses within the same GRB are confined to a rather narrow band of values. The wide range of values of $\\zeta$ are found not only in multi-peaked GRBs, but also in single-peaked GRBs that display smooth fast-rise, slow-decay light curves \\citep{br01,rp02}. The smooth single peak GRBs could arise from curvature effects \\citep{fmn96}, or to external shocks \\citep{dbc99}. For GRB pulses that could be produced by spherically symmetric shell collisions, discrepancy with observations suggest a breakdown of our assumptions. In the next section, the kinematic model is presented. Calculations based on this model are presented in Section 3. In Section 4, we discuss the possibility that radiative-cooling effects produce the power-law relation, implying a prediction that can be tested with Swift\\footnote{swift.gsfc.nasa.gov}. Alternately, the uniform spherical shell assumption could break down, or the basic model of colliding shells could be in error. The Appendices give derivations of simple, widely-used approximations related to this study, a derivation of the curvature relation $f_{\\epk} \\propto \\e_{pk}^3$, as well as an analytic form for the time-dependent pulse profile, leading to a simple expression for the light curve of a pulse in the curvature limit. A brief summary is given in Section 5. ", "conclusions": "The estimate $L \\cong 4\\pi d_L^2 \\Phi_E \\cong 4\\pi r_0^2 c u_0^\\prime \\Gamma^2$, where the received flux is intensified by two powers of $\\Gamma$ for the relativistic time contraction and photon energy enhancement in a blast wave geometry, is generally used to relate bolometric energy flux and internal energy density (Appendix A). More remarkably, the allowed radius of the radiating spherical shell is $\\approx 2 \\Gamma^2 $ times larger than inferred through causality arguments applied to the measured variability time scale. This effect greatly dilutes the comoving photon density compared with a stationary emitting region, and essentially explains the unusual properties of GRBs. In total, we see that \\begin{equation} \\Phi_{E} \\cong {cr_0^2 u_0^\\prime\\Gamma^2\\over d_L^2}= {4c^3 t_{var}^2\\over (1+z)^2 d_L^2}\\;u_0^\\prime \\Gamma^6\\; \\propto u_0^\\prime \\Gamma^6. \\label{phie1} \\end{equation} For the nominal parameters used in the figures, $\\Phi_E \\cong 4.8\\times 10^{-11} t_{var}^2 ({\\rm s})$ $u_0^\\prime \\Gamma_{300}^6$ ergs cm$^{-2}$ s$^{-1}$. The most rigorous limits on $\\gamma\\gamma$ attenuation are obtained by determining the {\\it minimum value} of the product $u_0^\\prime \\Delta \\tp$ that can produce a pulse with a measured peak flux $f_{\\e_{pk}}$ and full-width half-maximum duration $t_{1/2}$ for a given value of $\\Gamma$. It is the product $u_0^\\prime \\Delta \\tp$ that enters into the $\\gamma\\gamma$ attenuation (and photomeson) calculations. If the shell is found to be optically thick at some photon energy for a given value of $\\Gamma$ even in this case, then $\\Gamma$ must be larger if photons with the corresponding energies are detected. Inspection of the various cases shows that the pulse formed in the curvature limit produces the brightest measured flux and shortest duration for a given value of the product $u_0^\\prime \\Delta \\tp$. This is because the measured duration is due entirely to curvature effects, and the radiated energy is compressed into the shortest duration and brightest pulse in this limit. From the results of Appendices C and D, this implies that the expression \\begin{equation} u_0^\\prime \\Delta \\tp \\cong {[2^{1/(3-a)}-1](1+z) d_L^2 f_{\\e_{pk}} \\over 8 c^3 \\Gamma^5 t_{1/2}} \\label{u0cdt} \\end{equation} gives the smallest possible values for $u_0^\\prime c \\Delta \\tp$, and this expression will therefore yield the most reliable minimum Lorentz factors for $\\gamma\\gamma$ attenuation calculations derived from {\\it GLAST} or ground-based air Cherenkov telescope observations. The corresponding expression for the comoving photon spectral energy density is thus \\begin{equation} u^\\prime_{\\e^\\prime} \\cong {0.26(1+z) d_L^2 f_{\\e_{pk}} \\over 8 c^3 \\Gamma^5 t_{1/2} \\Delta \\tp} [x^a H(1-x)+x^b H(x-1)]\\;, \\label{ueprime} \\end{equation} where $t_{1/2}$ is determined at photon energies near the peak of the $\\nu F_\\nu$ spectrum. Eq.\\ (\\ref{ueprime}) is a factor of 3 smaller than the expression used for a comoving spherical blob with $\\delta \\rightarrow \\Gamma$ and $t_{var} \\rightarrow t_{1/2}$ (see Eq.\\ [2] in \\citet{der04}). Three generic types of pulses have been identified for the simple kinematic pulse, namely the curvature case where $r_0 \\gg c\\Gamma\\Delta \\tp$, the causal case where $r_0\\approx\\Gamma\\Delta r^\\prime \\approx c\\Gamma\\Delta \\tp$, and the thin shell case where $\\Delta r^\\prime \\ll c\\Delta \\tp$. In all three types of kinematic pulses, curvature effects dominate the formation of the spectrum at late time $t \\gg (1+z) \\Delta \\tp/2\\Gamma$, so that $f_{\\e_{pk}} \\propto \\epk^3$ if curvature effects dominate pulse formation at late times. The curvature relationship can be derived from a simple scaling argument by noting that the differential stationary-frame shell volume which contributes to the received flux, given by $dV =2\\pi r_0^2 \\Delta r d\\mu$, remains constant with time. This is because the relation between reception time $t$ and $\\mu$ for a shell that is instantaneously illuminated at comoving time $\\tp_0$ is $t = (1+z)\\Gamma\\tp_0(1-\\beta\\mu)$, so that $d\\mu \\propto dt$. The $\\nu F_\\nu$ flux $f_\\e =\\delta^4 L^\\prime/4\\pi d_L^2 = \\delta^4 V^\\prime \\ep j(\\ep )/4\\pi d_L^2 = \\delta^3 V \\ep j(\\ep )/4\\pi d_L^2$, where $L^\\prime$ is the comoving luminosity of the emitting volume that contributes to the flux at time $t$. For an emission spectrum that is flat, that is, $ \\ep j(\\ep) \\propto \\e^{\\prime 0}$, $f_{\\e_{pk}} \\propto \\epk^3$ because $\\epk \\propto \\delta$ for a uniform shell. Analysis of BATSE GRB light curves (Borgonovo and Ryde 2001) shows that the peak fluxes of a GRB pulse generally follow a relation whereby \\begin{equation} f_{\\e_{pk}}\\propto \\epk^\\zeta\\;. \\label{fepk} \\end{equation} Values of $\\zeta$ for different GRBs vary over a wide range from $\\approx 0.6$ to $3$, with values of $\\zeta$ roughly constant for pulses within the same GRB or in a GRB consisting of a single smooth pulse. In most GRBs, therefore, curvature effects do not make a large contribution to the decay phase of a GRB light curve. An interesting question is the source of the difference of observations from our kinematic model pulses. One possibility is that the jet has angular structure, and varies with directional energy release and baryon-loading on angles $\\theta \\approx $ few$\\times \\Gamma^{-1}$. The angle-dependent speeds in such a system would produce a deformed colliding shocked fluid shell where the spherical symmetry assumption fails, as therefore would the uniform jet model. If this is the case, then GRB prompt emission data can in principle be analyzed to reveal shell structure and to determine whether this behavior is consistent with a universal jet structure (\\citet{zha04}; see \\citet{fra03} for review). Rather than treat these geometrical effects here, we consider instead whether radiation effects could form a power-law relation between $f_{\\e_{pk}}$ and $ \\epk$. The most naive system considers a fixed volume of shocked fluid within which the peak of the $\\nu F_\\nu$ spectrum is made by a large population of quasi-monoenergetic electrons that radiates most of the power through the synchrotron process in a mean magnetic field of strength $B$. If these electrons mainly have comoving Lorentz factors $\\gamma$, then their luminous power $\\propto B^2\\gamma^2$. Because the peak of the $\\nu F_\\nu$ spectrum is $\\propto B \\gamma^2$, and assuming $B$ is constant, then $ f_{\\e_{pk}}\\propto \\epk$ or $ \\zeta_{syn} = 1$ for this simple synchrotron model with constant magnetic field. This model can therefore only apply in rare cases. A better treatment must consider the evolution of $\\gamma$ due to synchrotron and adiabatic losses in the expanding shell. The equation of electron energy evolution is given by \\begin{equation} -{d\\gamma\\over d\\tp} = {1\\over V^\\prime_{sh}} {dV^\\prime_{sh}\\over d\\tp}\\; {\\gamma\\over 3} + {\\sigma_{\\rm T}B^2(\\tp )\\over 6\\pi m_ec}\\;\\gamma^2\\;, \\label{dgammadt} \\end{equation} where the comoving shell volume changes with time according to $V^\\prime_{sh} \\propto t^{\\prime 3m}$, with $m = 0$ corresponding to no expansion, and $m = 1$ corresponding to 3-dimensional expansion. The magnetic field will also change as a result of the expansion of the shell volume. In the flux freezing limit where the magnetic field is randomly oriented, $BR^2 \\propto const$, implying $B \\propto V_{sh}^{\\prime 2/3}\\propto t^{\\prime -2 m}$. The well-ordered magnetic field required to explain the polarization observation of the GRB 021206 observed with RHESSI \\citep{cb03} suggests that there is not an efficient mixing and randomization of the magnetic field directions. For simplicity, we therefore write $B \\propto t^{\\prime -2 v m}$, where $v = 1$ gives the flux-freezing limit. Eq.\\ (\\ref{dgammadt}) becomes \\begin{equation} -{d\\gamma\\over d\\tau} = m\\; {\\gamma\\over \\tau} + \\nu_0\\tau ^{-4vm}\\gamma^2\\;, \\label{dgammadt1} \\end{equation} where $\\tau \\geq 1$ is a dimensionless time variable, and $\\nu_0$ is a dimensionless synchrotron energy loss rate. Eq.\\ (\\ref{dgammadt1}) is analytic, but it is sufficient to consider two limiting cases of dominant adiabatic losses or dominant synchrotron losses at late times. In the case of dominant adiabatic losses we have (dropping the primes) $\\gamma\\propto t^{ -m}$, $B\\propto t^{ -2 v m}$, and $\\e_{pk} \\propto t^{-2m (1+v)}$, so that $ f_{\\e_{pk}}/ \\epk \\propto B \\propto \\e_{pk}^{v/(v+1)}$. Thus $\\zeta_{adi} = 1 + [v/(v+1)]$. Even for a wide range of values of $v$, $1 \\lesssim \\zeta_{adi} \\lesssim 2$, and $\\zeta_{adi}$ is independent of the geometry factor $m$. If synchrotron losses dominate the cooling, $-d\\gamma/dt \\propto B^2 \\gamma^2$. The dependence $B\\propto t^{ -2 v m}$ therefore implies $\\gamma \\propto t^{4vm-1}$, so that $\\e_{pk} \\propto B\\gamma^2 \\propto t^{6vm-2}$. Hence $f_{\\e_{pk}}/\\e_{pk} \\propto B \\propto \\e_{pk}^{vm/(1-3vm)}$, so that $\\zeta_{syn}= 1+ [vm/(1-3vm)]$. Except when $v \\ll 1$, $\\zeta_{syn} \\approx 1$ when $m=0$ and $\\zeta_{syn} \\approx 0.5$ -- 0.67 when $m = 1$. In this simple model, therefore, values of $\\zeta$ between $1/2$ and $2/3$ are only possible for 3-dimensional expansion. Three-dimensional expansion is more likely to occur for narrowly collimated blast waves than for blast waves with large opening angles, and the narrowly collimated blast waves would have ``beaming breaks\" in the optical afterglow light curves at earlier times. If our conjecture that the $f_{\\e_{pk}}$/$\\e_{pk}$ relationships are due to synchrotron and adiabatic effects in GRB blast waves with different opening angles, then those blast waves with $\\zeta < 1 $ should be correlated with earlier beaming break times. Because this effect is only seen when synchrotron losses dominate the cooling, GRBs with $\\zeta < 1$ should also display cooling spectra with photon indices $\\approx 3/2$ below $\\e_{pk}$. \\citet{br01} find several GRBs and many pulses in the BATSE sample with statistically significant values of $\\zeta $ less than unity. These GRBs however preceded the afterglow era. For those GRBs which have measured beaming breaks (see Table 1 in \\citet{bfk03}), only GRB 990123 has sufficiently bright BATSE data to provide a data point for such a correlation. GRB 990123 has not yet been analyzed to give $\\zeta$, while analysis of Beppo-SAX data is in progress (F.\\ Ryde, private communication, 2004). Such a model for the $f_{\\e_{pk}}$/$\\e_{pk}$ relationship would explain why $\\zeta$ is approximately constant for different pulses within a GRB, provided that the opening angle of the GRB jet remains the same throughout the period of activity of the GRB engine. The adiabatic/synchrotron model would not, however, explain pulses with $2 \\lesssim \\zeta \\lesssim 3$. There are many such pulses in the \\citet{br01} sample, though generally with large error bars. If analysis of Beppo-SAX or Swift data reveal such GRBs, then another explanation is required. One possibility is that GRB pulses are due to the interactions of a single impulsive blast wave with inhomogeneities in the surrounding medium. This version of the external shock model for the prompt phase can be much more efficient than an internal shell model \\citep{dm99,dm03}, and permits quantitative studies of the statistics of BATSE GRBs (\\citet{bd00}; see \\citet{zm04} for a review of the internal/external controversy). Predictions for the $f_{\\e_{pk}}/\\e_{pk}$ relationship in an external shock model \\citep{dcb99} can be derived by adapting the equations for blast wave deceleration in a uniform medium with $\\Gamma = \\Gamma_0/[1+(x/x_d)^g]$, where $\\Gamma_0$ is the initial Lorentz factor, $x_d$ is the deceleration distance and $g$ is the radiative index ($g = 3/2$ and 3 for an adiabatic and fully radiative blast wave, respectively). In the deceleration phase, $x \\propto t^{1/(2g+1)}$ and therefore $\\Gamma \\propto t^{-g/(2g+1)}$. In this model, $\\e_{pk} \\propto \\Gamma B \\gamma_{pk}^2$ and $f_{\\e_{pk}} \\propto \\Gamma^2 B^2 \\gamma_{pk}^2$, where $\\gamma_{pk} \\propto \\Gamma^4\\propto t^{-4g/(2g+1)}$ in the slow-cooling regime, and $\\gamma_{pk} \\propto (x\\Gamma)^{-1}\\propto t^{-2/(2g+1)}$ in the fast-cooling regime. Thus $f_{\\e_{pk}}$/$\\e_{pk} \\propto B\\Gamma$. In the slow-cooling regime, $\\e_{pk} \\propto \\Gamma^4\\propto t^{-4g/(2g+1)}$, and $f_{\\e_{pk}}\\propto\\e_{pk}^{3/2}$. In the fast-cooling regime, $\\e_{pk} \\propto t^{-2/(2g+1)}$ and $f_{\\e_{pk}}\\propto\\e_{pk}^{1+g}$. In the slow-cooling and fast-cooling regimes, therefore, values of $\\zeta_{sc} = 3/2$ and $\\zeta_{fc} = 1+g$, respectively, are predicted. Provided that the surrounding medium is uniform (which can be inferred from afterglow modeling, though at a larger distance scale), the slow-cooling result implies a definite value of $\\zeta_{sc}=3/2$ for fast-rise, smooth decay light curves when spectral analysis demonstrates that the GRB evolves in the slow-cooling regime. For GRBs in fast-cooling regime, this estimate implies $5/2 < \\zeta_{fc} < 4$, and in these cases cooling spectra should be apparent. Further work will be needed to extend the results to radial density gradients of the circumburst medium, and to verify that these relations hold for deceleration in small density inhomogeneities that form GRB pulses in the external shock model." }, "0403/astro-ph0403214_arXiv.txt": { "abstract": "{We present the first {\\it BeppoSAX} observation (0.1 to 220 keV) of the quasar \\object{Mrk~205}. We have searched for the unusual Fe line profile claimed in the {\\it XMM-Newton} spectrum which has been widely discussed in recent literature. We find no evidence for a broad, ionized Fe line component in our data. We detect for the first time a Compton hump in this object. Besides, when this component is included in the fit, the line strength diminishes, in agreement with a recent re-analysis of the {\\it XMM-Newton} data, but with better constraints on the reflection component thanks to the PDS instrument (15-220 keV). We interpret this fact as another indication for illumination of a distant and cold material rather than reprocessing in the highly ionized inner parts of an accretion disk. \\\\ We cannot constrain the presence of a high energy cutoff but we confirm the existence of a variable soft excess (one year timescale). ", "introduction": "Discovered by Weedman (\\cite{weedman}), the quasar \\object{Mrk~205} is seen through the southern spiral arm of the nearby spiral, barred galaxy \\object{NGC~4319} at a projected distance of roughly 40\\arcsec\\, from its center. From its radio luminosity at 6 cm, \\object{Mrk~205} is classified as a radio-quiet quasar (Rush et al. \\cite{rush}), but it is still not clear whether it belongs to the radio-quiet quasar class (e.g. Bahcall et al. \\cite{bahcall}) or to the Seyfert~1 class (e.g. Veron-Cetty \\& Veron \\cite{veron}). \\object{Mrk~205} has a redshift $z=0.07085$ (Huchra et al. \\cite{huchra}) while \\object{NGC~4319} has $z=0.00468$ (Bowen et al. \\cite{bowen1}). \\\\ \\\\ In the standard paradigm of (radio-quiet) Active Galactic Nuclei (AGN), X-ray photons result from Compton upscattering of optical-UV photons in a hot thermal corona above the accretion disk surrounding a supermassive black hole. A large fraction of the seed photons are thought to be produced by thermal emission of the accretion disk itself, but \\object{Mrk~205} was found by McDowell et al. (\\cite{mcdowell}) to have a weak blue bump. One of the possible scenarios which could explain this spectral morphology is that the bump may be highly variable and thus observed in a particularly weak state. In the thermal comptonisation framework, blue bump variability would also induce hard X-ray variations.\\\\ A soft X-ray excess component may be interpreted as the hard tail of the blue bump (Walter \\& Fink \\cite{walterfink}; Brunner et al. \\cite{brunner}). So far, evidence for a soft excess in \\object{Mrk~205} was only reported in two observations; with {\\it EINSTEIN} IPC (Wilkes \\& Elvis \\cite{wilkeselvis}) and {\\it XMM-Newton} (Reeves et al. \\cite{reeves2}).\\\\ {\\it BeppoSAX}, with its unique broad-band capabilities allows to extend the observation of \\object{Mrk~205} for the first time above 20 keV to study the properties of the hard X-ray emission as well as of a potential reflection component. Indeed, some of the X-ray photons can be reprocessed in the surroundings, in particular in the disk itself giving rise to neutral (or low-ionization) Fe~K$\\alpha$ line. While it is a common feature of Seyfert~1 galaxies (Nandra \\& Pounds \\cite{nandra}), only a few quasars present a clear detection of such a component (Williams et al. \\cite{williams}; Nandra et al. \\cite{nandra2}). Furthermore, the centroid of the line was often measured at energies close to 6.7 keV suggesting that its origin may be in an ionized layer of the accretion disk (Reeves \\& Turner \\cite{reeves1}), the neutral Fe~K$\\alpha$ being emitted in distant and cold material lying outside the broad line region. Reeves et al. (\\cite{reeves2}), who analysed the {\\it XMM-Newton} data of \\object{Mrk~205}, claimed the existence of a broad ionized component in addition to the neutral one in this object. However, Page et al. (\\cite{page}) reprocessed the data including a reflection hump component in the fit and found little evidence for a broad, ionised component in the data. \\\\ In this paper, we present the analysis of three {\\it BeppoSAX} observations of \\object{Mrk~205} for a total of 200 ks of exposure time. In Section 2, we summarize the knowledge we had on \\object{Mrk~205} prior to our {\\it BeppoSAX} observation. In Section 3, we describe the data and their analysis while in Section 4 we compare our results and discuss them in the framework of the unified model. ", "conclusions": "\\noindent {\\bf The nature of \\object{Mrk~205}.} It has been noted that, in contrast to the Seyfert~1 case, the detection of Fe K$\\alpha$ line in quasars was more seldom and that if detected, the line energy was around 6.7 keV instead of 6.4 keV (Reeves \\& Turner \\cite{reeves1}). It was thus suggested that sensitive instruments like the EPIC camera on {\\it XMM-Newton} could observe the Fe K$\\alpha$ emission of quasar in detail. From our X-ray observations of \\object{Mrk~205}, and from the luminosity we obtained from them, we suggest to be careful in classifying this object. Its luminosity is found in the region in which it is difficult to choose for a Seyfert~1 or for a quasar. It seems thus difficult to generalize any findings made on this object to a whole class. \\\\ \\\\\\noindent {\\bf The soft excess variability.} The soft excess seems to be variable in this source although a very strong soft excess was never reported. It seems to be absent/not detected in 1997, strong in 2000, weak in 2001 and absent again in 2002 (thus varied in a 2 weeks timescale). The variability of the soft excess cannot be associated with variability of a cold absorber as we have shown in Fig.~\\ref{Fig1} and Section 3 that the absorbing column density deduced from spectral modeling was roughly constant over all the observations. \\\\ Note that the soft excess component measured here should not be associated with direct thermal emission from the accretion disk because its temperature is much too high. \\\\ \\\\\\noindent {\\bf Thermal comptonisation.} The historic variations of the photon index can be explained in the framework of thermal comptonisation. \\object{Mrk~205} showed large $2-10$ keV flux variations between the 1997 (high flux state), 2000 (low flux state), 2001 (high flux state) and 2002 (low flux state) observations which do not seem to be correlated with variability of the soft excess. \\\\ However, in thermal comptonisation processes, the index of the comptonised component is inversely correlated with the ratio of the photon seed flux to the comptonised power law flux. We can test this hypothesis if we assume that the soft excess strength is proportional to the strength of the soft photon field (Walter \\& Fink \\cite{walterfink}; Walter \\& Courvoisier \\cite{walter1}). The index for 2000 was $1.73\\pm 0.02$ for a ratio of 23\\% while for 2001 the index was $2.00\\pm 0.03$ for a ratio of 3\\%. This is thus in disagreement with our expectations. \\\\ On the other hand, using the {\\sc compha} model to estimate the UV flux, we found that an increase of the photon index goes with a larger UV flux, in agreement with simple thermal comptonization model expectation. Similar conclusions were found by Petrucci et al. (\\cite{pet00}) for the Seyfert~1 galaxy \\object{NGC~5548}. The contradiction between the soft X-ray and UV variations is not very surprising. Indeed the soft excess component is certainly linked to the UV bump through complex radiative transfer (e.g. comptonization in the warm (few keV) upper layer of the accretion disk). Consequently, its expected spectral behavior is not straightforward and may be at odds from the zeroth order expectations. \\\\ \\\\\\noindent {\\bf Constraints on the reflecting material.} The nature of the reprocessing medium can be the accretion disk or some obscuring, distant material. If the Fe K$\\alpha$ line is produced in the inner parts of the accretion disk, its profile should be considerably broader, with very distinct asymmetries due to Doppler and gravitational redshift. This is not possible to test with the {\\it BeppoSAX} data while Page et al. (\\cite{page}) report that they did not find evidence for it in the {\\it XMM-Newton} data either.\\\\ The strength of the Compton reflection $R$ is here systematically higher than what was found in Seyfert galaxies ($R\\simeq 0.8\\pm 0.2$, Nandra \\& Pounds \\cite{nandra}), while on the contrary, the iron line equivalent width EW is smaller than the mean EW value. It is worth noting however that our measures of $R$ are consistent with $R<1$ at 90\\% confidence, see Fig.~\\ref{Fig10}.\\\\ The inconsistency between the best fit values of $R$ and the line EW may then occur by chance and the reflecting material in \\object{Mrk~205} may simply cover a solid angle $<2\\pi$ as seen from the X-ray source. The line energy then also suggests the material to be neutral. A low Fe abundance will make the line weaker and the Compton reflection hump stronger (as observed) by reducing the opacity above the Fe K edge (George \\& Fabian \\cite{geofab}).\\\\ Given these constraints a simple solution could be that the reprocessing is done in a distant material e.g. the dust torus.\\\\ \\\\ However, we cannot rule out the presence of close mildly ionized material. Indeed Nayakshin \\& Kallman (\\cite{naykal}) have shown that in the case of relatively strong illumination by local X-ray flares, the upper layer of the reflecting material may become strongly ionized. In this case the reflection features are mainly produced by the neutral depth layer and the expected line energy is 6.4 keV as we observed. The line EW is however smaller than the one expected in the case of neutral reflector since the line may be partly comptonized in the hot skin. But clearly this skin cannot be highly ionized since in this case it would act like a perfect mirror and the reflection hump would be suppressed contrary to what we observe. Lines in the soft band are also expected in the case of a mildly ionized medium and it may explain the poor fit of the soft excess with a simple black body component. \\noindent \\\\ \\\\ {\\bf The 1 keV line.} The spectrum of January 2002 contains a narrow emission feature centered at an energy of $\\sim$ 1 keV which was also found in at least 3 other AGN. This feature could be explained by a blend of several emission lines, predominantly from ionized species of Ne and possibly from Fe L shell. The line emission could be a result of reflection in photoionized matter. It could also be resulting from recombination in an optically thin thermal plasma." }, "0403/astro-ph0403200_arXiv.txt": { "abstract": "We consider the general conditions of quark droplets formation in high density neutron matter. The growth of the quark bubble (assumed to contain a sufficiently large number of particles) can be described by means of a Fokker-Planck equation. The dynamics of the nucleation essentially depends on the physical properties of the medium it takes place. The conditions for quark bubble formation are analyzed within the frameworks of both dissipative and non-dissipative (with zero bulk and shear viscosity coefficients) approaches. The conversion time of the neutron star to a quark star is obtained as a function of the equation of state of the neutron matter and of the microscopic parameters of the quark nuclei. As an application of the obtained formalism we analyze the first order phase transition from neutron matter to quark matter in rapidly rotating neutron stars cores, triggered by the gravitational energy released during the spinning down of the neutron star. The endothermic conversion process, via gravitational energy absorption, could take place, in a very short time interval, of the order of few tens seconds, in a class of dense compact objects, with very high magnetic fields, called magnetars. ", "introduction": "Since \\citet{Wi84}, following early proposals by \\citet{It70} and \\citet{Bo71}, suggested that strange quark matter, consisting of $% u$-, $d$- and $s$-quarks is energetically the most favorable state of the matter, the problem of existence of strange quark stars has been intensively investigated in the physical and astrophysical literature. The possibility that some compact objects could be strange stars remains an interesting and intriguing, but still open question. \\citet{Wi84} also proposed two ways of formation of strange matter: the quark-hadron phase transition in the early universe and conversion of neutron stars into strange ones at ultrahigh densities. In the theories of strong interaction quark bag models suppose that breaking of physical vacuum takes place inside hadrons. As a result vacuum energy densities inside and outside a hadron become essentially different, and the vacuum pressure on the bag wall equilibrates the pressure of quarks, thus stabilizing the system. If the hypothesis of the quark matter is true, then some of neutrons stars could actually be strange stars, built entirely of strange matter \\citep{AlFaOl86,HaZdSch86}). However, there are general arguments against the existence of strange stars \\citep{CaFr91}. For an extensive review of strange star properties see \\citet{ChDaLu98}. Several mechanisms have been proposed for the formation of quark stars. Quark stars are expected to form during the collapse of the core of a massive star after the supernova explosion, as a result of a first or second order phase transition, resulting in deconfined quark matter \\citep{Da}. The proto-neutron star core, or the neutron star core, is a favorable environment for the conversion of ordinary matter to strange quark matter \\citep{ChDa}. Another possibility is that some neutron stars in low-mass X-ray binaries can accrete sufficient mass to undergo a phase transition to become strange stars \\citep{Ch96}. This mechanism has also been proposed as a source of radiation emission for cosmological $\\gamma $-ray bursts \\citep{Ch98a}. Some basic properties of strange stars like mass, radius, cooling, collapse and surface radiation have been also studied \\citep{ChHa,ChHa03}. The physical mechanisms of the transition from neutron matter to quark matter in an astrophysical background have been studied within several models. The first is due to \\citet{Ol87}, who used a non-relativistic diffusion model. As such, this is a slow combustion model, with the burning front propagating at a speed of approximately $10$ m/$\\sec $. This is determined primarily by the rate at which one of the down quarks inside the neutrons is converted, through a weak decay, to a strange quark: $% d+u\\rightarrow s+u$. The second method of describing the conversion process was first suggested by \\citet{HoBe88}, and analyzed in detail by \\citet{LuBeVu94} and \\citet{LuBe95}, who modelled the conversion as a detonation. In this case the conversion rate is several orders of magnitude faster than that predicted by the slow combustion model. This model is based on the relativistic shock waves and combustion theory. But regardless of the way in which the transformation occurs, an initial seed of quark matter is needed to start the process. Real neutron stars have two conserved charges- electric and baryonic. Therefore, a neutron star has more than one independent component and in this sense it is a ``complex\" system \\citep{Gl02}. The characteristics of a first-order phase transition, like the deconfinement transition, are very different in the two cases. In the ``complex\" system the conserved charges can be shared by the two phases in equilibrium in different concentrations in each phase. The mixed phase, formed from hadrons and quarks cannot exist in a simple body in the presence of the gravity because the pressure in that phase is a constant \\citep{Gl00}. This causes a discontinuity in the density distribution in the star occurring at the radius where Gibbs criteria are satisfied. The isospin symmetry energy in neutron rich matter will exploit the degree of freedom of readjusting the charges between hadronic and quark phases in equilibrium so as to reduce the symmetry energy to the extent consistent with charge conservation. Regions of hadronic matter will have a net positive charge neutralized by a net negative charge on the quark matter regions \\citep{Gl00}. Coulomb repulsion will prevent the regions of like charge to grow too large, and the surface energy will act in the opposite sense in preventing them to become too small. Consequently, the mixed phase will form a Coulomb lattice so as to minimize the sum of Coulomb and surface interface energy at each proportion of phases. As the quark phase becomes more abundant, the droplets merge to form rods, rods merge to form slabs etc. \\citep{Gl00}. Hence the actual geometric phase of quark nuggets in a neutron star evolves as a function of the pressure and of the surface energy between dense nuclear matter and quark matter. In order to describe the process of formation and evolution of quark nuclei in the neutron stars one must use nucleation theory. The goal of nucleation theory is to compute the probability that a bubble or droplet of the $A$ phase appears in a system initially in the $B$ phase near the critical temperature \\citep{LaLi80}. Homogeneous nucleation theory applies when the system is pure. Nucleation theory is applicable for first-order phase transitions when the matter is not dramatically supercooled or superheated. If substantial supercooling or superheating is present, or if the phase transition is second order, then the relevant dynamics is spinodal decomposition \\citep{ShMo01}. The nucleation of strange quark matter inside hot, dense nuclear matter was investigated, with the use of Zel'dovich's kinetic theory of nucleation, by \\citet{HoBeVu92}. By assuming that the newly formed strange quark matter bubble can be described by a simple bag model containing $N_{q}$ ultrarelativistic quarks, and by assuming that the time evolution of the radius of the bubble is described by an equation of the form $dr/dt=(r-R_{c})/\\tau _{w}$, where $R_{c}$ is the critical radius of the bubble and $\\tau _{w}$ the weak interaction time scale, the nucleation rate $\\xi $ is given by \\begin{equation} \\xi =2.2\\times 10^{-2}\\left( T/\\sigma \\right) ^{1/2}N_{qc}^{3/4}\\tau _{w}^{-1}\\exp \\left( -3.1N_{qc}^{1/2}\\sigma /T\\right) , \\end{equation} where $N_{qc}\\leq 300$ is the critical quark number, $\\sigma $ is the surface tension of the bubble and $n_{n}$ and $n_{q}$ are the particle number densities in the neutron and quark phase, respectively \\citep{HoBeVu92}% . In this way a lower bound for the temperature of the nucleation, $T\\geq 2.1MeV$, can be obtained. The effects of the curvature energy term on thermal strange quark matter nucleation have been considered by \\citet{Ho94}, who derived, within the same approach, the following expression for the nucleation rate: \\begin{equation} \\xi =\\frac{r_{c}^{2}n_{q}n_{n}}{4\\pi \\tau _{w}}\\left( T/\\sigma \\right) ^{1/2}\\exp \\left( -4\\pi R_{c}^{2}/3T\\right) . \\end{equation} All the effects of the curvature enter only through the value of $R_{c}$. As a general conclusion one obtains the result that even the curvature term acts against strange quark matter nucleation, the physical temperature of a just born proto-neutron star is, in any model, more than enough to drive an efficient boiling of the neutron material (the observations of the neutrino flux from SN1987A are consistent with an effective temperature of $T\\sim 4$ MeV). The possibility of stable strange quark matter (both bulk and quasi-bulk) at finite temperature, and some of its properties have been investigated, within the framework of the dynamical density-dependent quark mass model of confinement, by \\citet{Ch93}. The behavior of the surface tension and the stability of quark droplets at $T\\neq 0$ has also been discussed, with reference to strangelet formation in ultra-relativistic heavy-ion collisions. A different approach to the problem of thermal nucleation was taken by \\citet{OlMa94}. They used the assumption (standard in the theory of bubble nucleation in first-order phase transitions) that bubbles form at a rate given by $R\\approx T^{4}\\exp \\left( -W_{c}/T\\right) $, with $% W_{c}$ the minimum work required to form a bubble with radius $r_{c}$. As a main result it follows that if the bag constant lies in the interval where three-flavor but not two-flavor quark matter is stable at zero pressure and temperature ($145$ MeV$\\leq B^{1/4}\\leq 163$ MeV), then all or parts of a neutron star will be converted into strange matter, during the first second of its existence. For bag constant above the stability interval, a partial transformation is still possible \\citep{OlMa94}. \\citet{He95} calculated the rate of formation of quark matter droplets in neutron stars from a combination of bubble formation rates in cold degenerate and high temperature matter, taking into account nuclear matter calculations of the viscosity and thermal conductivity. The droplet formation rate is that of \\citet{LaTu73}, given by $% I=\\left( k/2\\pi \\right) \\Omega _{0}\\exp \\left( -W_{c}/T\\right) $, with $% \\Omega _{0}=\\left( 2/3\\sqrt{3}\\right) \\left( \\sigma /T\\right) ^{3/2}\\left( r_{c}/\\xi _{q}\\right) ^{4}$ the statistical prefactor, which measures the phase-space volume of the saddle point around $R_{c}$ that the droplet has to pass, in its way to the lower energy state. $\\xi _{q}$ is the quark correlation length. The dynamical prefactor determines the droplet growth rate and is given by $k=\\left( 2\\sigma /\\Delta w^{2}r_{c}^{3}\\right) \\left( \\lambda T+2\\left( 4\\eta /3+\\zeta \\right) \\right) $, where $\\Delta w$ is the enthalpy difference and $\\lambda ,\\eta $ and $\\zeta $ are the thermal conductivity and the shear and bulk viscosities, respectively. The droplet formation rate can be expressed as \\citep{He95} \\begin{equation}\\label{eqn} I=\\frac{\\sigma _{20}^{7/2}\\mu _{400}^{2}\\eta _{50}}{\\Delta P_{10}\\Delta w_{10}^{2}T_{10}^{3/2}}\\exp \\left( 185-134\\frac{\\sigma _{20}^{3}}{\\Delta P_{10}^{2}T_{10}}\\right) \\textrm{ s}^{-1}\\textrm{km}^{-3}, \\end{equation} with $\\mu _{400}$ is the quark chemical potential in units of $400$ MeV. The total number $N$ of droplets formed by nucleation is the integrated rate over the volume of the neutron star and the time after the nucleation process starts at the moment $t_0$, $N=\\int_{0}^{R}4\\pi r^{2}dr\\int_{t_{0}}^{\\infty }I\\left[ \\Delta p\\left( r\\right) ,T\\left( t\\right) \\right]dt$, where $R$ is the radius of the neutron star. The pressure and temperature depend sensitively on the equation of state of the nuclear and quark matter. To convert the core of the neutron star into quark matter at least one droplet must be formed, i.e., $N>1$. Then, with the use of Eq. (\\ref{eqn}), the condition for the formation of at least one droplet requires $\\sigma \\leq 24$ MeV fm$^{-2}\\Delta P_{c,10}^{2/3}T_{c,10}^{1/3}$, where $P_{c}$ and $T_{c}$ are the core values of the pressure and temperature, respectively \\citep{He95}. The pressure and enthalpy difference depends strongly on the equation of state of nuclear and quark matter and hence the droplet formation rate cannot be reliably estimated. The effect of subcritical hadron bubbles on an inhomogeneous first order quark-hadron phase transition was studied in \\citet{ShMoGuGl00}. The transition from nuclear matter (consisting of neutrons, protons and electrons) to matter containing strangeness was considered, within a mean field type model and with the use of Langer nucleation theory, by \\citet{No02}. An estimate of the time the new phase to appear at various densities and times in the cooling history of a protoneutron star was also given. On the other hand, as a result of an increase of the central density of the star, due for example to accretion or spinning down, a metastable, super-compressed neutron phase can appear, with a star consisting of the new phase surrounded by normal neutron matter \\citep{Gr98}. This situation is similar to a gas undergoing a phase transition to its liquid state if compressed to a volume $V_{c}$ at fixed $T$. If it is slowly compressed, it may stay in the vapor state even for $V 10$ and, subsequently, fragmented into stars due to molecular hydrogen cooling \\citep{loeb01}. These collapsing objects then fragmented into many clumps, which had typical masses of $\\sim 10^{\\,2}-10^{\\,3}M_{\\odot }$. Very massive stars have lifetimes of $\\sim 3\\times 10^{\\,6}\\,\\rm{years}$ and end their lives as supernovae. \\par Recently, Lanzetta et al. \\citep{lanzetta02} showed that the incidence of the highest intensity star formation regions increases monotonically with redshift. Their observations indicate that star formation in the early universe occurred at a much higher rate than was previously believed. Therefore, the rate of occurrence of supernovae would have also been much higher in the past than at the present. Supernovae shocks disturb the plasma in which they are immersed, producing turbulent motions of the gas. If the supernovae rate was much higher in the past than at present, the plasma of the first formed objects must have been much more turbulent than that of presently observed, low redshift, star forming galactic molecular clouds. \\par Turbulence generates stochastic magnetic fields (magnetic noise) at a faster rate than it does mean fields \\citep{kul-and-92}. If the turbulence is strongly non-helical, the fields induced are confined to small scales (Kazantzev \\citealp{kazantzev68}). However, if it is helical, induction of large scale magnetic correlations by the $\\alpha$ effect occurs (Vainshtein \\& Kichatinov \\citealp{vainshtein86}). Astrophysical turbulence is mainly of a helical nature. Hence, we can expect that large scale correlations were induced by the high redshift, turbulent plasma. \\par In this study, we explore the hypothesis that the magnetic fields observed in high redshift galaxies were created by small scale, stochastic, turbulent helical dynamos, rather than by mean field dynamos. We use a simple model of a gas cloud that is assumed to have collapsed at a high redshift $z > 10.$ At $z \\sim 10,$ the cloud would have had a low magnetization level and a high level of turbulence. Thus, it would have been similar to the turbulent, low ionization, star forming molecular clouds observed in our galaxy, albeit with a much smaller initial magnetic field and a much higher turbulence level due to the higher star formation rate in the early universe. \\par It is well known that shock waves produced by supernova explosions accelerate cosmic rays to energies $\\geq\\,\\rm{GeV}$ (e.g., \\citealp{cr-acc}). V\\\"{o}lk et al. \\citep{volk89} showed that in all galaxies, the supernova rate is a direct measure of the cosmic ray intensity. We can, therefore, infer that cosmic rays were already present in considerable intensities in high redshift galaxies. We take into account phenomenologically, the effect on turbulence of cosmic rays, supernova shocks, and powerful stellar winds from massive stars on turbulence by varying the turbulent parameters over a broad range, in order to take into account the uncertainies in our knowledge of high redshift structures. \\par The linear evolution equations for the correlation function of magnetic fields for non-helical turbulence were derived nearly forty years ago by Kazantzev \\citep{kazantzev68}. For helical turbulence, the corresponding equations were obtained twenty years later by Vainshtein \\& Kichatinov \\citep{vainshtein86}. These equations are linear in the magnetic correlations. Recently, Subramanian \\citep{subramanian99} and Brandemburg \\& Subramanian \\citep{bran-sub} derived the non-linear evolution equations for the magnetic correlations by taking into account the back-reaction of the Lorentz force on the plasma charges in the form of ambipolar drift. We solve the nonlinear helical evolution equations numerically for various values of the parameters that characterize the high redshift turbulent plasma. \\par The paper is organized as follows. In section II, we give the evolution equations for the magnetic correlations. We describe the effects of the main parameters on the integration in section III . Finally, in section IV, we summarize and discuss our results. ", "conclusions": "In this work, we studied the problem of the origin of strong coherent large-scale magnetic fields, observed in low-redshift galaxies and previously thought to have been created by the mean field dynamo. Since doubts have been cast on the mean field dynamo as the source of these fields due to their observation in high redshift objects, where the dynamo would not have had suficient time to operate, we investigated here the stochastic helical dynamo as such a source. We showed that these fields can be generated in the plasmas found in very high redshift objects. \\par The generation of strong large scale coherent fields is a highly non-linear magnetohydrodynamical problem, which depends upon many factors. Here we discussed a possible mechanism for the generation of large scale fields, namely the non-linear evolution and diffusion of magnetic noise. Magnetic noise becomes coherent on a scale which is larger than that of the turbulence due to the presence of non linear terms in the evolution equations for the magnetic correlations. \\par We found that for realistic turbulent parameters, it is possible to generate the magnetic fields, observed at high redshifts . Between $z\\sim 10$ and $z\\sim 5$, the primordial plasma is strongly turbulent and partially ionized The magnetic field intensities reached are independent of initial correlations. \\par We considered a very simple model for the generation and evolution of the magnetic correlations. The dependence of the resulting magnetic field intensity on the charge composition of the primordial plasma, suggest that the reionization and star formation processes played an important role in determining the features of the magnetic fields detected in high redshift objects. In a forthcoming work we shall address the evolution of the magnetic correlations, considering a time dependent ion density as well as other nonlinear processes, such as the the Hall effect. Other turbulent scenarios, in addition to the homogeneous and isotropic one considered here, will be treated as well." }, "0403/astro-ph0403170_arXiv.txt": { "abstract": "{ We study the cooling of isolated neutron stars. The main cooling regulators are introduced: EoS, thermal transport, heat capacity, neutrino and photon emissivity, superfluid nucleon gaps. Neutrino emissivity includes main processes. A strong impact of medium effects on the cooling rates is demonstrated. With taking into account of medium effects in reaction rates and in nucleon superfluid gaps modern experimental data can be well explained. ", "introduction": "\\label{sec:intro} The Einstein Observatory was the first that started the experimental study of surface temperatures of isolated neutron stars (NS). Upper limits for some sources have been found. Then ROSAT offered first detections of surface temperatures. Next $X$-ray data came from Chandra and XMM/Newton. Appropriate references to the modern data can be found in recent works by \\cite{TTTTT02,T04,KYG01,YGKLP03}, devoted to the analysis of the new data. More upper limits and detections are expected from satellites planned to be sent in the nearest future. In general, the data can be separated in three groups. Some data show very {\\em{``slow cooling''}} of objects, other demonstrate an {\\em{``intermediate cooling''}} and some show very {\\em{``rapid cooling''}}. Now we are at the position to carefully compare the data with existing cooling calculations. ", "conclusions": "We have shown that the most up-to-date observed NS cooling data can well be explained. Actually we deal with a many-parametric problem that allows for a variation of many quantities. Besides the data points might be partially shifted from their positions shown in figures due to a number of uncertainties and assumptions used at their analysis. By above figures we have illustrated different possibilities discriminating more probable explanations from less probable ones. We elaborated the example of the probably most realistic and microscopically supported EoS of the $V18+\\delta v +UIX^*$. Actually we used a simple parameterization of this EoS suggested by \\cite{HJ99} (HHJ). In this EoS the DU process does not show up to $M\\simeq 1.839 M_{\\odot}$. Thus within this model the {\\em ``Standard + DU''} scenario would demonstrate that the majority of experimentally measured cooling points relate to very massive NS, with $M\\geq 1.84 ~M_{\\odot}$. From our point of view such a scenario seems unrealistic. We exploit in-medium effects in the calculation of the emissivities and the pairing gaps, and, that is less important, in specific heat and heat conductivity. In general, medium effects result in a significant suppression of the superfluid gaps, especially of the $3P_2$ $nn$ gap, and they enhance the cooling rates of the MMU and the MNB processes through {\\em the pion softening effect with the increase of the density.} Without the latter effect the {\\em ``Standard + PU''} scenario suffers from an internal inconsistency. Pion condensation cannot take place without preliminary softening of the pion mode at lower densities. And v.s., recent argumentation for the pion condensation, cf. \\cite{APR98,SST99}, further motivates the presence of a precursor pion softening, details see in \\cite{MSTV90}. Pion condensation at $n\\gsim 3n_0$ does not contradict the data but the data can be explained also without pion condensation and other so called \"exotics\" (KU, HDU, etc) but with the pion softening. Based on the works \\cite{SVSWW97,BGV01} we have further improved our code. We showed that the results are sensitive to the values and the density dependencies of the nucleon superfluid gaps. {\\em The strong suppression of the $3P_2$ gap} motivated by theoretical evaluations that incorporate medium effects is indeed required for the fit of the data. We show that all three groups of points {\\em ``slow cooling''}, {\\em ``intermediate cooling''} and {\\em ``rapid cooling''} are now well explained on the basis of the {\\em ``Nuclear medium cooling scenario'' demonstrated here, where an important r\\^ole is played by in-medium effects}, cf. \\cite{V00}. We may draw the following main conclusions. i) The normal matter assumption (see Figs. \\ref{fig6} - \\ref{fig10}) seems rather unrealistic, as by itself, as in relation to the cooling data. One could explain the data but at the price of ignoring of medium effects in MMU and MNB. Then the {\\em ``intermediate cooling''} points can only be explained by very low NS masses (as $0.5~M_{\\odot}$) or, together with {\\em ``rapid cooling''} points, by the very high NS masses $M>1.839~M_{\\odot}$, which allow for the DU process. In the latter case the mass window separating the {\\em ``intermediate cooling''} and {\\em ``rapid cooling''} is very narrow. Above price seems us too high and we drop such a scenario. The superfluidity is called for by the data. ii) Including superfluid gaps we see, in agreement with recent microscopic findings, that $3P_2$ neutron gap should be as small as $\\lsim 10$~keV. Otherwise one cannot explain at least several old objects (see Figs. \\ref{fig19}, \\ref{fig19a}). Proton and $1S_0$ neutron gaps also might be suppressed by factors of order of several, as it is demonstrated by Figs. \\ref{fig20}, \\ref{fig21}, \\ref{fig15}, \\ref{fig15a}, but the suppression by factor $\\gsim 10$ is not already permitted. iii) Medium effects associated with the pion softening are called for by the data. As the result of the pion softening the pion condensation may occur for $n\\geq n_c^{\\rm PU}$ ($n\\geq 3n_0$ in our model). Its appearance does not contradict to the data (see Fig. \\ref{fig16}) but also the data are well described, if the softening effect is rather saturated (see Fig. \\ref{fig12}) with increase of the density (as demonstrated by curves 1a, 1b in Fig. \\ref{fig1}). At the same time the critical densities for the efficient DU-like exotic processes (such as PU on charged pion) should not be as small as $<2.5~n_0$, if the given HHJ EoS is indeed correct. Otherwise one would get too rapid cooling of the NS of the typical $1.4~M_{\\odot}$ mass. This also means that the proper DU threshold density can't be too low that puts restrictions on the density dependence of the symmetry energy. Both statements might be important in the discussion of the heavy ion collision experiments. iv) We demonstrated a regular mass dependence: for the NS masses $M\\gsim 1.0~M_{\\odot}$ less massive NS cool slower, more massive NS cool faster. As we have mentioned, for the sake of simplicity we did not include the possibility of the hyperonization and other possibilities, like kaon condensation and fermion condensation, which may stimulate a more rapid cooling, being working in a line with the pion condensation. We did not include possible quark effects. The latter need a special treatment. The possibility of the color superconductivity in dense NS interiors opens a number interesting possibilities like the so called two-flavor color superconductivity (2SC) phase, color-flavor-locking (CFL) phase, color-spin-locking (CSL) phase. Their cooling is essentially different. We will return to this discussion in the nearest future." }, "0403/astro-ph0403616_arXiv.txt": { "abstract": "{ The currently available values and confidence limits for $\\Omega_{\\mathrm{m0}}$, $H_0$ and globular cluster ages still indicate that the dark energy that dominates the Universe could also be a form of quintessence or phantom energy. In fact, current cosmological values favor phantom energy. To increase the likelihood of a cosmological constant as dark energy instead of phantom energy, the possibilities seem to lie in reducing globular cluster ages, the Hubble constant, or both, and possibly advancing the epoch of globular cluster formation. For a set of possible dark energy equations of state that includes the cosmological constant, quintessence or phantom energy, age--redshift analytical expressions for null curvature universes that include ordinary matter are derived together with the corresponding ages for these universes. ``Cosmic coincidences'' are found for the currently accepted values of $\\Omega_{\\mathrm{m0}}$ when evaluating the age of the Universe with a cosmological constant or a phantom energy as dark energies. ", "introduction": "Curvature, age, the Hubble constant and energy density form a long-standing cosmological puzzle. Measurements of the cosmic microwave background radiation (CMBR) anisotropy have confirmed that the Universe is consistent with null, or nearly null, curvature (de Bernardis et al.\\ \\cite{bernardis}), as predicted by the inflation paradigm. However, a Euclidean universe composed of ordinary matter only has an age inconsistent with that derived from globular clusters (Bolte \\& Hogan \\cite{bolte}). Moreover, galaxy cluster data (Carlberg et al.\\ \\cite{carlberg}) allow us to obtain only about 20--30\\% of the amount of matter required for achieving null curvature. This puzzle seemed to be solved by the discovery in 1998 of the first evidence for an accelerating universe based on distant type Ia Supernova (SNeIa) observations (Riess et al.\\ \\cite{riess1}), which have been supported by additional data including more distant SNe (Riess et al.\\ \\cite{riess2}). This acceleration implies the presence of a significant fraction of dark energy with an equation of state of negative pressure. The addition of the required amount of this energy to the observed amount of matter allows us to obtain the total amount of energy required for achieving null curvature. It also has the desired side effect of producing an older universe that can then be made consistent with independent age measurements. However, the nature of this dark energy is still to be established. Moreover, its equation of state is still unknown. The first attempts rescued the cosmological constant as an ad hoc explanation, although their interpretation of the constant in terms of a vacuum energy is inconsistent by 124 orders of magnitude with respect to the required value, which leads us to assume that its value has to be zero for consistency. Other energy types, such as quintessence or phantom energy, have also been invoked, although their origin is no clearer than that of the cosmological constant. Quintessences were introduced by Wetterich (\\cite{wetterich}), Caldwell et al. (\\cite{caldwell1}), Ratra \\& Peebles (\\cite{ratra}) to avoid the extreme fine-tuning needed to allow a cosmological constant to be significant only at recent epochs. Quintessences are characterized by a scalar field whose evolution depends on its potential. The most popular potentials are inverse power and exponential laws (Kneller \\& Strigari \\cite{kneller} and references therein), although other possibilities have been considered (see, for example, Di Pietro \\& Claeskens\\ \\cite{dpc}). Both inverse power laws and exponential potentials converge to a unique solution for a broad range of different initial conditions. Phanton energy (Caldwell \\cite{caldwell2}) violates the dominant-energy condition (Hawking \\& Ellis\\ \\cite{he}) that might allow the existence of wormholes. It also makes the Universe reach in a finite time a cosmic doomsday where all objects, from galaxies to nucleons, are ripped apart, a situation that has been termed the ``big rip'' (Caldwell \\cite{caldwell2}). Although some Chaplygin gas model generalization can avoid this big rip (Gonz\\'alez-D\\'\\i az \\cite{gd}), the violation of the dominant energy condition and the difficulties in obtaining a stable phantom model (Carroll et al.\\ \\cite{carroll}) renders this kind of energy more problematic than quintessences. In the case of either quintessences, the cosmological constant or phantom energy, it seems established that the introduction of negative pressure energies is required to fit and harmonize the existing CMBR anisotropy, SNeIa, clusters of galaxies, large scale structure, Big Bang nucleosynthesis and age estimator data. The incorporation of these energies changes the age--redshift relation and estimates of the age of the Universe. The aim of this article is to show the difficulty in avoiding phantom energies with the current age estimator data and to provide analytical age expressions for the homogeneous and isotropic case for the most likely equations of state for our Universe, considered constant, as restricted by the cosmological parameters commonly assumed. These include the cosmological constant, quintessence and phantom energy. Thomas \\& Kantowski (\\cite{thomas}) provide an age--redshift relation using elliptical integrals for a universe composed of matter and a cosmological constant, but without providing an easy ready-to-use expression. In general, no analytical expressions of age as a function of redshift seems to be readily available in the literature for the currently most widely accepted cosmologies: flat universes of matter plus a cosmological constant, quintessence or phantom energy. Although numerical integration allows us to obtain ages in a relatively straightforward way (see, for example, Hogg \\cite{hogg}), analytical solutions are faster to evaluate and have the advantage of providing more precise and ready results than numerical approaches. Moreover, analytical expressions provide the explicit dependence on energy densities, thereby easing their study. Finally, they constitute a reference for checking numerical solutions. In Section 2 the basic equations are given. In Section 3 possible values for the equation of state, considered constant, are reviewed according to recent figures for the cosmological parameters, in Section 4 the general case of a non-constant equation of state is outlined, and in Section 5 analytical age--redshift relations and ages are obtained by solving the equations on a case by case basis for the range of the constant equations of state assumed. ", "conclusions": "In the most accepted cosmological model the Universe is currently accelerated owing to a for{\\bf m} of dark energy of unknown origin, as found via distance--redshift fits to high redshift SNeIa. This energy complements the fraction of ordinary matter so that the curvature is null, according to the observed CMBR anisotropy and the inflation paradigm. Also, the dark energy increases the age of the Universe for the same values of $\\Omega_{\\mathrm{m0}}$ and $H_0$ with respect to a universe composed of non-relativistic matter only, thus alleviating possible age conflicts. In this article a recent determination of $\\Omega_{\\mathrm{m0}}$, $t_0$ and $H_0$ has been selected from the literature with the criterion of providing a set of cosmological parameters as model-independent as possible. A simple analysis shows that the weighted mean equation of state of dark energy has upper values in the range $-0.23 > w > -1.21$. Hence, not only a cosmological constant but also quintessence or phantom energy are also viable dark energy candidates from an observational point of view. In fact, the mean values of $\\Omega_{\\mathrm{m0}}$, $t_0$ and $H_0$ favor phantom energy over other alternatives. Moreover, a time-dependent $w$ favors even more the existence of epochs with $w<-1$. It is qualitatively shown that to reduce the likelihood of phantom energy to favor a cosmological constant as dark energy, smaller globular cluster ages and/or smaller $H_0$ values with respect to those currently determined are required. Pushing back globular cluster formation to earlier epochs cannot be ruled out, but only together with lower values of $t_{\\mathrm{GC}}$ and/or $H_0$. Otherwise, an age conflict might still be present. The range of upper $w$ limits inferred is approximately covered by taking $w=-n/3$ for $n=1,2,3,4$. For each of these constant values of $w$, which include quintessence, the cosmological constant and phantom energy, analytical age--redshift solutions for Euclidean universes have been deduced and analytical age expressions derived. Some of these analytical solutions have not been derived before. Furthermore, no analytical expressions seem to be available in the literature for the range of the constant equations of state considered. Analytical equations are more precise, faster and readier to use than numerical calculations when evaluating the age of the Universe. Moreover, the explicit dependence on $\\Omega_{\\mathrm{m0}}$ and $\\Omega_{\\mathrm{w0}}$ renders their study easier. Finally, curious ``cosmic coincidences'' make 1/$H_0$ a good approximation for the age of the Universe for the currently accepted $\\Omega_{\\mathrm{m0}}$ values assuming null curvature and a cosmological constant or a phantom energy as dark energies." }, "0403/astro-ph0403420_arXiv.txt": { "abstract": "We present detailed comparisons between high quality observational colour-magnitude diagrams (CMDs) of open star clusters and synthetic CMDs based on MonteCarlo numerical simulations. The comparisons account for all of the main parameters which determine the shape of the CMD for a stellar population. For the four clusters studied, NGC 6819, NGC 2099 (M37), NGC 2168 (M35) and NGC 2323 (M50), we derive reddening, distance, age, binary fraction, star formation rate and indicative metallicity by comparing the locations and density of points in the observed CMDs to the simulated CMDs. We estimate the uncertainties related to stellar evolution theories by adopting various sets of stellar models for all of the synthetic CMDs and discuss which stellar models provide the theoretical CMDs that best reproduce the observations. ", "introduction": "Theoretical isochrones are commonly fit to the major observational sequences of star clusters in order to both better understand the underlying physics of stellar evolution and to determine properties of the clusters, i.e., the age. If the metallicity, reddening and distance to the cluster are well constrained from independent techniques, the comparisons typically involve matching the morphology of the turn-off and location of the red giant stars to predictions. Recently, the newer method of using synthetic colour-magnitude diagrams to compare with observational data has proven to be much more informative and rewarding \\citep{aparicio,tosi,skillman}. These MonteCarlo simulations allow modelling of several additional parameters which dictate the distribution of points in the CMD, such as stochastic star formation (SF) processes, binary fraction, photometric spread, main-sequence thickness, data incompleteness and small number statistics. Consequently, the results not only provide a measure of the properties of the cluster, but can also constrain the star formation history (SFH) and the initial mass function (IMF). Furthermore, by comparing the simulations based on several different sets of evolutionary tracks, we can constrain which models use the best prescription of parameters (such as treatment of overshooting, mixing length, etc...). Confronting the simulations with observations requires a large data set with accurate photometry. For this, we use the deep $BV$ photometry presented in the CFHT Open Star Cluster Survey (\\cite{kalirai1}, hereafter JSKI). JSKI observed 19 open star clusters in our Galaxy and have yet published results on the four richest clusters, NGC 6819 (\\cite{kalirai2}, hereafter JSKII), NGC 2099 (\\cite{kalirai3}, hereafter JSKIII), and NGC 2168 and NGC 2323 (\\cite{kalirai4}, hereafter JSKIV). These data were reduced and calibrated in a homogenous manner as described in JSKI. The resulting CMDs exhibit very tight main sequences showing several `kinks' and slope changes which are predicted by theory. More importantly, the combination of very short and deep exposures, and the large aerial coverage of the detector (42$' \\times$ 28$'$) has allowed the measurement of stars from the brightest asymptotic giant branch (AGB) and red giant branch (RGB) phases down to very low-mass main-sequence phases ($\\sim$0.2 M$_\\odot$). This allows our comparisons to yield evolutionary information over a wide mass range. The reduced data set in the CFHT Open Star Cluster Survey has been requested by, and made available to, several investigators for additional science rewards outside our goals (e.g., astrometric studies, proper motions, variable stars, radial velocities, brown dwarfs, blue stragglers and Galactic disk star distributions). The present study complements these efforts and analyses the four published clusters in a way that allows us to include them in a large homogeneous sample of open clusters aimed at studying the formation and evolution of the Galactic disk (Bragaglia 2003, and references therein). Galactic open clusters are indeed particularly well suited to this purpose, since they span a range of ages from a few million to several billion years and can be observed in various regions of the Galactic disk characterised by different star formation histories. They can be used to study both the present day disk structure and its temporal evolution (Janes \\& Phelps 1994, Friel 1995, Tosi 2000, Bellazzini et al. 2003). Old open clusters offer a unique opportunity to trace the whole kinematical and chemical history of our disk, if collected in populous and representative samples and accurately and homogeneously analysed (see e.g., Twarog, Ashman, \\& Anthony-Twarog 1997; Carraro, Ng, \\& Portinari 1998). Here we apply the synthetic CMD method to NGC 6819, NGC 2099, NGC 2168 and NGC 2323 to derive their age, reddening, distance modulus and (approximate) metallicity homogeneously to the Bragaglia (2003) cluster sample. The method also allows us to determine other features of these clusters, such as the existence (or lack thereof) of a significant fraction of unresolved binary systems, the original total mass of formed stars and the possible evaporation of some of the lower mass stars. The organisation of the paper is as follows, \\S \\ref{observations} briefly summarises the data and the reduction procedures. Further details are given in JSKI. In \\S \\ref{firstresults} we present a summary of our main results which relate to this work from the published papers in the CFHT Open Star Cluster Survey. \\S \\ref{synthetic} sets up the numerical simulations and presents details on how the synthetic CMDs were created. Next, we compare the CMDs and the corresponding luminosity functions from the observations with the simulations on a cluster-by-cluster basis (\\S \\ref{vs}). Finally, we discuss the results in \\S \\ref{discussion} and conclude the study in \\S \\ref{conclusions}. ", "conclusions": "\\label{conclusions} We re-derive key parameters for the four very rich open star clusters, NGC 6819, NGC 2099, NGC 2168, and NGC 2323. The parameters are measured by comparing high quality, deep empirical colour-magnitude diagrams with MonteCarlo simulations. The combination of comparing the morphology and number of stars in various evolutionary phases and cluster luminosity functions allows us to provide tight constraints on the reddening, distance, age and binary fraction for each cluster. In cases where the cluster metallicity is not certain, simulations with different abundances are also compared. In all cases the data are better reproduced when a fraction of unresolved binary systems between 20 and 30\\% is assumed. A summary of the results is given in Table 2. The synthetic CMDs and LFs are generally found to be in excellent agreement with the observational data. This circumstance, combined with the fact that different sets of stellar evolution tracks provide different values for the cluster parameters, confirms how important it is to use more than one set of models to estimate the theoretical uncertainties. It also shows that a homogeneous approach is crucial to derive reliable overall cluster properties, such as age-metallicity relations. In fact, cluster dating based on different stellar models may lead not only to different absolute ages, but also to different age ranking, with significant drawbacks on the interpretation of the cluster properties in terms of Galactic evolution. Some discrepancies, such as the thickness of the main sequences {\\it vis \\`a vis} the size of the photometric errors, and the possibility that cluster stars may be contaminating our blank field (particularly in the case of NGC 2168) are identified and discussed. For each cluster, we also measure the astration mass (i.e., the total mass that went in all of the stars formed in the cluster) according to a single slope Salpeter IMF. From this, we calculate the star formation rate between 0.1--100 $M_{\\odot}$ to be $9.4 \\times 10^{-6}, 1.3 \\times 10^{-5}, 6.2 \\times 10^{-6}$, and $2.7 \\times 10^{-6} M_{\\odot}yr^{-1}pc^{-2}$ for NGC 6819, NGC 2099, NGC 2168, and NGC 2323 respectively. These rates would be a factor 1.6 lower if the IMF below 0.6 $M_{\\odot}$ has a slope of +0.44, as inferred by Gould et al. (1997) from HST data, rather than Salpeter. The true value may in fact lie somewhere in between these two IMFs, as recently discussed e.g., by Chabrier (2003). In principle our data are deep enough to allow for a direct derivation of the cluster IMF; however, the four clusters are heavily contaminated by fore/background stars and the lack of appropriate decontamination fields has prevented us from a safe analysis of the star counts at the fainter magnitudes. The need of appropriate photometry in nearby fields is emphasised, also for the purpose of adequate studies of the clusters mass segregation/evaporation." }, "0403/astro-ph0403566_arXiv.txt": { "abstract": "We place observational constraints on a recently proposed Galactic population, dubbed the {\\it shroud} (Gyuk \\& Gates 1999, Gates \\& Gyuk 2001). The shroud would be a very thick Galactic disk of low luminosity objects, most likely old white dwarfs, proposed to explain the optical depth seen in microlensing surveys towards the Magellanic clouds. The shroud is a simple alternative to the lenses being distributed in a classical, near-spherical dark halo; the advantage of the shroud is that it would compose only a fraction of a dark halo's total mass. In this paper, we argue that stars of the Galactic shroud would be detectable in the recent proper motion survey of Oppenheimer et al. (2001) if their absolute luminosities were brighter than $M_{R_{59F}} = 19.4$ or approximately $M_V = 18.6$. We adopt a range of simple models of the shroud's kinematics and morphology, and the colours and luminosities of its white dwarfs; via Monte-Carlo simulations, we predict the numbers expected in the Oppenheimer et al. survey, which would be clearly separated from the numbers produced by white dwarfs of the disk, thick disk and halo. The number of white dwarf detections in the proper motion survey (98) is found to be well explained by the disk, thick disk and halo. With {\\it the most conservative} kinematic and density parameters for the shroud, and an absolute luminosity of the white dwarfs of $M_{R_{59F}} = 17.6$, we find that the proper motion survey would detect over 100 WDs, just from the shroud. For a $M_{R_{59F}} = 19.4$ shroud, the survey would find $5 \\pm 2$ peculiar objects, whereas only two white dwarfs with such characteristics are found in the original data. $M_{R_{59F}} = 19.4$ corresponds to $M_V = 18.6$ for WDs with $(V - I) = -1.030$. ", "introduction": "The microlensing surveys carried out in recent years (e.g. EROS, MACHO and OGLE) have reported on lensing population of dark objects seen towards the Magellanic clouds. The favoured mass for these objects is approximately half a solar mass, suggesting they are white dwarfs, since main sequence M stars of this mass are bright enough to be detected directly in surveys. To date, no directly detected counterpart for the dark population has been found. One scenario is a population of ancient white dwarfs (WDs) which would comprise a significant fraction of the Galactic dark halo (up to 20 per cent of its mass). This suggestion has later become disfavoured because even quite dim WDs would be directly detectable in the most recent proper motion surveys even if they comprised `only' 2 per cent of the total mass of the dark halo (Reyl\\'e, Robin \\& Cr\\'ez\\'e 2001; Flynn, Holopainen \\& Holmberg 2003). To explain the MACHO microlensing results (Alcock et al. 2000), and at the same time to avoid some of the problems with a massive dark halo population, Gates and Gyuk (hereafter, G\\&G) (1999, 2001) proposed a new population in the Milky Way, dubbed the Galactic `shroud'. It would be a WD population in the form of a very thick disk with a scale height of 2.0 -- 3.0 kpc. It would produce the same microlensing optical depth as a dark halo WD population without having to be enormously massive. We study here the implications of the shroud using the same techniques that we used in a previous study (Flynn et al. 2003; hereafter, Paper I). In Paper I, we constrained the luminosity and the number density of a dark halo WD population with the same simulation that we use for the current study. In the previous study, we used two proper motion surveys independently for constraining the halo population. After testing the simulation for Paper I, we are now confident to use only the more recent one of those surveys (Oppenheimer et al. 2001) for constraining the new population. Oppenheimer et al. proper motion survey was originally designed to find dark halo WDs. However, the follow up studies that have investigated the possibility of dark halo WDs in this survey have found that the survey is also very sensitive to the conventional thick disk WDs (e.g., Paper I; Reid, Sahu \\& Hawley 2001; Reyl\\'e et al. 2001). Because the shroud has a similar velocity structure to the thick disk, the survey is also very sensitive to the shroud. Thus, the Oppenheimer et al. survey is optimal for our purposes. In section 2, we briefly introduce how to find nearby WDs and separate them by stellar population. In section 3, we go through the parameters of the model in detail, and in section 4, we describe the Oppenheimer et al. proper motion survey and its findings. We illustrate the effect of proper motion to the predicted number counts in section 5 and present our results in section 6. Finally, we conclude in section 7. ", "conclusions": "Strong limits have been placed on the luminosities of white dwarfs which could make up the putative `shroud' of the Galaxy, proposed as a solution to the optical depth measurements seen in the microlensing surveys towards the Magellanic clouds. We use the Oppenheimer et al. (2001) proper motion survey of 4000 square degrees to $R_{59F} = 19.7$, containing 98 spectroscopically confirmed WDs. A range of shroud models are investigated, and the number of WDs with high reduced proper motions compared to the Oppenheimer et al. data. Most of the models produce significantly more WDs than are actually observed; in particular, models in which we probe the very highest reduced proper motion source (indicating very low luminosity and high space velocities, as expected for the shroud component) allow us to limit the luminosity of the WDs in the shroud to $M_{R_{59F}} = 19.4$ (the survey pass band), which corresponds to $M_V = 18.6$ or $M_I = 19.6$. If the Galaxy possesses a shroud of WDs which produces the microlensing signal, these WDs must be fainter than the above limits. Very few models of Hydrogen atmosphere WDs cool to such faint levels within the age of the Universe." }, "0403/astro-ph0403085_arXiv.txt": { "abstract": "\\cbstart The Chandra X-ray Observatory grating spectrometers allow study of stellar spectra at resolutions on the order of 1000. Prior x-ray observatories' low resolution data have shown that nearly all classes of stars emit x-rays. Chandra reveals details of line and continuum contributions to the spectra which can be interpreted through application of plasma models based on atomic databases. For cool stars with hot coronae interpreted in the Solar paradigm, assumption of collisional ionization equilibrium allows derivation of temperature distributions and elemental abundances. Densities can be derived from He-like ion's metastable transition lines. Abundance trends are unlike the Sun, as are the very hot temperature distributions. For young stars, there is evidence of accretion driven x-ray emission, rather than magnetically confined plasma emission. For some hot stars, the expected emission mechanism of shocked winds has been challenged; there is now evidence for magnetically confined thermal plasmas. The helium-like line emission in hot stars is susceptible to photoexcitation, which can also be exploited to derive wind structure. \\cbend ", "introduction": "X-ray emission is ubiquitous among late-type and pre-main sequence stars, as has been amply demonstrated by imaging and low-resolution x-ray observatories.\\citep{Feigelson:Montmerle:1999} With the advent of the Chandra transmission-grating and the XMM-Newton reflection-grating spectrometers, we can now probe the nature of the x-ray emission in detail through high-resolution diagnostics. Early Chandra results confirmed some of the abundance anomalies derived from low-resolution imaging spectra and also unambiguously confirmed that few-component temperature models generally are not adequate. Chandra spectroscopy has also challenged long-standing hot-star wind and x-ray production theories. Much effort is now being spent to survey and analyze stellar x-ray spectra over a range of evolutionary states, spectral types, rotational periods, and activity levels. Here we will examine some results for a variety of stars, the ``active'' binaries; young, low-mass stars; and hot, high mass stars, with emphasis on Chandra grating spectrometers. We will not discuss results from the XMM-Newton observatory, though they are complementary in many ways. ", "conclusions": "" }, "0403/hep-ph0403115_arXiv.txt": { "abstract": "An analysis is carried out within mSUGRA of the estimated number of events originating from upward moving ultra-high energy neutralinos passing through Earth's crust that could be detected by the Extreme Universe Space Observatory (EUSO). The analysis exploits a recently proposed technique that differentiates ultra-high energy neutralinos from ultra-high energy neutrinos using their different absorption lengths in the Earth's crust. It is shown that for the part of the parameter space, where the neutralino is mostly a Bino and with squark mass $\\sim 1$ TeV, EUSO could see ultra-high energy neutralino events within mSUGRA models with essentially no background. In the energy range $10^{9}~{\\rm GeV} < E_{\\tilde \\chi} < 10^{11}~{\\rm GeV}$ the unprecedented aperture of EUSO makes the telescope sensitive, after 3~yr of observation, to neutralino fluxes as low as $d\\Phi/dE_{\\tilde \\chi} > 1.1 \\times 10^{-6}\\ (E_{\\tilde \\chi}/{\\rm GeV})^{-1.3}$~GeV$^{-1}$ cm$^{-2}$\\ yr$^{-1}$\\ sr$^{-1},$ at the 95\\% CL. Such a hard spectrum is characteristic of supermassive particles' $N$-body hadronic decay. The case in which the flux of ultra-high energy neutralinos is produced via decay of metastable heavy ($m_X = 2 \\times 10^{12}$~GeV) particles with uniform distribution throughout the universe, and primary decay mode into 5 quarks + 5 squarks, is analyzed in detail. The normalization of the ratio of the relics' density to their lifetime has been fixed so that the baryon flux produced in the supermassive particle decays contributes to about 1/3 of the events reported by the AGASA Collaboration below $10^{11}$~GeV, and hence the associated GeV $\\gamma$-ray flux is in complete agreement with EGRET data. For this particular case, EUSO will collect between 4 and 5 neutralino events (with 0.3 of background) in $\\approx 3$~yr of running. NASA's planned mission, the Orbiting Wide-angle Light-collectors (OWL), is also briefly discussed in this context. ", "introduction": "mSUGRA~\\cite{msugra} and its extensions (generically called SUGRA models) are currently the leading candidates for physics beyond the standard model. These models contain a consistent mechanism for the breaking of supersymmetry softly by gravity mediation. An attractive feature of these models is that with R parity conservation the lightest neutralino is a possible candidate for cold dark matter~\\cite{goldberg} in a significant part of the mSUGRA parameter space~\\cite{scaling}. Further, over most of the parameter space the phenomenon of scaling occurs~\\cite{scaling} so that the light neutralino is mostly the supersymmetric partner of the $U(1)_Y$ gauge boson $B_{\\mu}$, i.e., it is mostly a $U(1)_Y$ gaugino or a Bino~\\cite{scaling,roberts}. The parameter space of mSUGRA is characterized by the universal scalar mass, $m_0$, the universal gaugino mass, $m_{\\frac{1}{2}}$, the universal trilinear coupling, $A_0$ (all taken at the grand unification scale $M_G\\sim 2\\times 10^{16}$ GeV), $\\tan\\beta =/$ where $H_2$ gives mass to the up quark, and $H_1$ gives mass to the down quark and the lepton. In addition the model contains the Higgs mixing parameter $\\mu$ which enters in the superpotential in the form $\\mu H_1H_2$. The magnitude of $\\mu$ is determined by the constraint of radiative electro-weak symmetry breaking in the theory while $sign\\mu$ is arbitrary and must be constrained by experiment. mSUGRA has been put to stringent test by the recent precision data from the satellite experiment, the Wilkinson Microwave Anisotropy Probe (WMAP) which imposes a narrow range for cold dark matter (CDM) so that~\\cite{bennett,spergel} $ \\Omega_{\\rm CDM} h^2 =0.1126^{+0.008}_{-0.009}$. The candidacy of neutralinos as the dark matter of the universe is based on relic densities surviving annihilation processes of non-relativistic particle. Detailed analyses show that mSUGRA allows for a small amount of the parameter space in agreement with WMAP observations~\\cite{wmap1}. As a consequence, the applicability of mSUGRA demands that the contribution of other sources of CDM to the dark matter mix are negligible. In this paper, we will be interested in a flux of ultrarelativistic neutralinos resulting from decays of a population of CDM metastable superheavy particles~\\cite{Berezinsky:1997hy,Kuzmin:1997cm}. In concert with the previous statement, these particles should contribute negligibly to the dark matter density. The weak couplings of neutralinos imply an interaction length in air which is greater than the atmospheric depth, even at horizontal incidence. The interaction probability is then roughly uniform throughout the atmosphere. As with neutrinos, showers initiated by neutralino primaries can be distinguished from hadronic events by restricting the zenith angle space to near horizontal -- this maximizes the probability to detect showers of weakly interacting primaries, while screening out the electromagnetic component of hadronic showers which are initiated high in the atmosphere. However, deeply developing neutralino cascades cannot be isolated from neutrino induced air showers. In this paper we show that the part of the parameter space where the neutralino is mostly a Bino and the mass $m_{\\tilde q}$ of the first and second generation squarks is $\\sim$ 1 TeV can lead to ultra-high energy neutralino signals that may be seen by the Extreme Universe Space Observatory (EUSO)~\\cite{Catalano:mm,Scarsi:fy}. These two conditions are fully compatible with the WMAP constraint, and the neutralino as lightest supersymmetric particle. Further, with appropriate cuts the background events arising from ultra-high energy neutrinos are essentially negligible. We discuss now the details of the analysis. The problem of discriminating between neutralino and neutrino induced showers with space-based experiments has been examined recently~\\cite{Barbot:2002et}. The method makes use of the Earth as a filter. Specifically, in the region of the mSUGRA parameter space under consideration the cross section for neutralino-nucleon interaction is smaller than that for neutrino-nucleon scattering processes. Thus, by restricting the angular bin for arrival of upward going showers to a region where neutrinos are largely absorbed during traversal in the Earth, it may be possible to obtain a clean signal~\\cite{Anchordoqui:2001cg}. In Ref.~\\cite{Barbot:2002et}, the discussion was presented in terms of neutralino-nucleon cross sections parameterized as a series of constant fractions of the neutrino-nucleon cross section. In this paper, we first calculate the neutralino cross section in the squark-resonance approximation. We then proceed to estimate the sensitivity of EUSO to neutralino-induced air showers. The sensitivity will be characterized by a lower bound on the neutralino flux, which is then related to some particular models of $X$-particle decay. ", "conclusions": "Using a technique that exploits the different absorption lengths of neutrinos and neutralinos in the Earth's crust, we have estimated the sensitivity of EUSO to isolate upward coming showers of ultra-high energy $\\tilde\\chi$. The neutralino-nucleon interaction has been approximated by resonant squark production, with the neutralino being largely Bino in composition, and $m_{\\tilde q}\\simeq$ 1 TeV. We have shown that, during the complete mission lifetime, the telescope will be sensitive to $E_{\\tilde \\chi}^2\\ d\\Phi/dE_{\\tilde \\chi} > 1.1 \\times 10^{-6}\\ (E_{\\tilde \\chi}/{\\rm GeV})^{0.7}$ GeV cm$^{-2}$\\ yr$^{-1}$ sr$^{-1}$ at the 95\\% CL, for $10^{9}~{\\rm GeV} < E_{\\tilde \\chi} < 10^{11}~{\\rm GeV},$ and for the region $m_{\\tilde q}=1.0\\pm 0.2$ TeV. A hard spectrum $\\propto E_{\\tilde\\chi}^{-1.3}$ is typical of super heavy relic $N$-body decays that are purely hadronic. This is a conservative estimate, since regeneration effects have been only considered in computing the neutrino background. We have explicitly analyzed the case in which the flux of ultra-high energy neutralinos is produced via decay of metastable heavy ($m_X = 2 \\times 10^{12}$~GeV) particles with uniform distribution throughout the universe, and primary decay mode into 5 quarks + 5 squarks. The normalization of $n_X/\\tau_X$ has been fixed to contribute about 1/3 of the events reported by the AGASA Collaboration below $10^{11}$~GeV~\\cite{Takeda:2002at}. For this particular case, EUSO will collect between 4 and 5 neutralino events (with 0.3 of background) in $\\approx 3$~yr of running. Existing limits on the diffuse photon flux in the GeV region strongly limit the sensitivity of EUSO for primary 2-body decays of hadronic nature. This is because, for the normalization of the baryonic contribution to the ultra-high energy cosmic ray flux assumed in Ref.~\\cite{Barbot:2002et}, which is marginally consistent with new EGRET bounds, the accompanying neutralino flux produced for $q \\bar q$ and $q \\tilde q$ is about an order of magnitude below the flux generated in the primary 10-body decay discussed above. On the other hand, the telescope can still be sensitive to the leptonic mode $X \\rightarrow l \\tilde l.$ In this case, by reducing $n_X/\\tau_X$ by a factor of $\\sim$ 2, one lessens the problem with EGRET data and still leaves a window open for neutralino detection at EUSO. We note that for a Bino-like neutralino, the primary decay mode (90\\% branching fraction) of the squark is $\\tilde q\\rightarrow q \\tilde g,$ with a subsequent decay $\\tilde g\\rightarrow q\\bar q \\ \\tilde\\chi.$ Thus, the neutralino energy of the decay is about 1/6 of the primary energy. In the remaining 10\\% of the decays, $\\tilde q\\rightarrow q \\tilde\\chi.$ In either case, the shower energies are far above the $\\sim$ 1 PeV threshold for the detector. If more detailed considerations are warranted in the future, regeneration effects during passage through Earth can be assessed, taking into account the energy losses of the decay modes. These effects will lead to some enhancement of $P$, and consequently of the event rate. We turn now to a brief discussion on the potential of the planned NASA mission Orbiting Wide-angle Light-collectors (OWL)~\\cite{Stecker:2000ek}. This mission will involve photo detectors mounted on 2 satellites in low equatorial orbit (600 - 1200 km). The eyes of the OWL will stereoscopically image a geometric area of $\\sim 9 \\times 10^5$~km$^{2},$ yielding $A \\sim 3 \\times 10^{6}$~km$^2$ sr. With its superior effective aperture ($\\epsilon_{_{\\rm DC}} \\approx 10\\%$), a 10 yr mission lifetime will allow one to discern on contributions of metastable relics to the upper end of the cosmic ray spectrum at the level of 1 part in $10^2.$ Consequently, the data from this mission will allow one to probe more deeply the parameter space of mSUGRA and its extensions." }, "0403/astro-ph0403017_arXiv.txt": { "abstract": "We present an analysis of the {\\em Chandra} High Energy Transmission Grating Spectrometer observation of the rapidly rotating ($P_{\\rm rot}=0.94\\,{\\rm d}$) post T~Tauri ($\\sim20 $\\,Myr old) star PZ~Telescopii, in the Tucana association. Using two different methods we have derived the coronal emission measure distribution, $em(T)$, and chemical abundances. The $em(T)$ peaks at $\\log T = 6.9$ and exhibits a significant emission measure at temperatures $\\log T > 7$. The coronal abundances are generally $\\sim 0.5$ times the solar photospheric values that are presumed fairly representative of the composition of the underlying star. A minimum in abundance is seen at a first ionization potential (FIP) of 7-8\\,eV, with evidence for higher abundances at both lower and higher FIP, similar to patterns seen in other active stars. From an analysis of the He-like triplet of \\ion{Mg}{11} we have estimated electron densities of $\\sim 10^{12}-10^{13}\\,{\\rm cm^{-3}}$. All the coronal properties found for PZ~Tel are much more similar to those of AB~Dor, which is slightly older than PZ~Tel, than to those of the younger T~Tauri star TW~Hya. These results support earlier conclusions that the soft X-ray emission of TW~Hya is likely dominated by accretion activity rather than by a magnetically-heated corona. Our results also suggest that the coronae of pre-main sequence stars rapidly become similar to those of older active main-sequence stars soon after the accretion stage has ended. ", "introduction": "\\label{intro} One of the primary characteristics of low mass pre-main sequence (PMS) stars is their intense X-ray activity. This X-ray emission therefore represents an important means for investigating the properties and evolution of young stellar objects. X-ray activity is present during the evolution of PMS stars both in the initial evolutionary stages of a Classic T~Tauri Star (CTTS, Class I and II sources), during which the star has an accretion disk that surrounds it, and in the subsequent Weak-Line T~Tauri Star (WTTS, Class III sources) phase in which the star has no accreting material and is approaching the zero-age main sequence \\citep[ZAMS, ][]{FeigelsonMontmerle1999}. How the presence of accreting material influences the X-ray emission of PMS stars, and for how long, remain questions of debate. The high resolution X-ray spectra now available with the {\\em Chandra} and {\\em XMM-Newton} satellites offer the possibility to perform detailed studies of stellar coronae because key emission lines diagnostics can now be resolved. These diagnostics can be used to derive elemental abundances, temperature and density structure of the emitting plasmas. It is worth noting that very few young stars in star forming regions or associations are sufficiently X-ray bright to allow high resolution X-ray spectroscopy with current instrumentation. In particular, among the CTTSs, \\objectname{TW~Hya} is the best studied case to date because it is the nearest ($\\sim56$\\,pc) known CTTS. TW~Hya shows spectral characteristics very different from those of young but otherwise \\emph{normal} active stars \\citep{KastnerHuenemoerder2002,StelzerSchmitt2004}: very low plasma temperature ($\\log T \\sim 6.5$), high density ($\\log N_{\\rm e} \\sim 13$), very low Fe abundance ($A_{\\rm Fe}/A_{\\rm Fe\\sun} \\sim 0.2$). \\citet{KastnerHuenemoerder2002} and \\citet{StelzerSchmitt2004} attributed these characteristics to an accretion shock rather than coronal activity. Without the benefit of high resolution spectra of similar stars for comparison, however, the nature of the peculiarity of TW~Hya remains uncertain. The situation with slightly more evolved stars is more clear. Studies of \\objectname{AB~Dor}, a young active star that has nearly arrived at the ZAMS and which has been observed in the past with several space-borne X-ray observatories, have shown that it is characterized by a hot corona ($T \\sim 10^7$\\,K) with plasma densities ranging from $6\\times10^{10}\\,{\\rm cm^{-3}}$ at $2\\times10^{6}$\\,K, to $3\\times10^{12}\\,{\\rm cm^{-3}}$ at $10^{7}$\\,K, and a moderately low Fe abundance ($A_{\\rm Fe}/A_{\\rm Fe\\sun} \\sim 0.25$, \\citealt{MeweKaastra1996,GudelAudard2001}; \\citealt*{Sanz-ForcadaMaggio2003}; P. Testa, in preparation; D. Garc{\\'{\\i}}a-Alvarez, in preparation). In many respects, AB~Dor can be considered the prototype of very active single stars. High resolution X-ray spectroscopy of other PMS stars with ages between few $10^{6}$\\,yr, typical of CTTSs like TW~Hya, and $10^{8}$\\,yr, the ZAMS of solar-type stars like AB~Dor, is crucial for understanding how the characteristics of stellar X-ray activity change during these early evolutionary phases. In this work, we present a {\\em Chandra} High Energy Transmission Grating Spectrometer (HETGS) observation of \\objectname{PZ~Telescopii} (HD~174429, HIP~92680). Classified as a K0V star by \\citet{Houk1978}, PZ~Tel was determined by \\citet{ZuckermanWebb2000} to be a member of the Tucana association, a nearby star forming region about 45\\,pc away. From observations with the ROSAT PSPC, \\citet{StelzerNeuhauser2000} deduced a PZ~Tel X-ray luminosity of $L_{\\rm X}\\sim(2.88\\pm0.0 8)\\times10^{30}\\,{\\rm erg\\,s^{-1}}$. At a Hipparcos parallax distance of $49.7\\pm2.9$\\,pc \\citep{PerrymanLindegren1997}, PZ~Tel is a single star with a rotational period of 0.94\\,d (\\citealt{CoatesHalprin1980}; \\citealt*{InnisCoates1984,InnisThompson1986}). \\citet{FavataMicela1998} have estimated a mass of $1.1\\,M_{\\sun}$ for PZ~Tel and an age of approximately 20\\,Myr. Its youth is also confirmed by prominent H$\\alpha$ emission and a relatively undepleted Li abundance \\citep*{SoderblomKing1998}. Further evidence of its PMS status has been pointed out by \\citet{BarnesCollierCameron2000} who deduced from $v \\sin i$ and $P_{\\rm rot}$ that the minimum radius of PZ~Tel is $R\\,\\sin i \\sim 1.27\\,R_{\\sun}$: this value is larger than the radius of a main sequence star with the same mass as PZ~Tel. The analysis of {\\em Chandra} HETGS high resolution spectra of PZ~Tel offers us the opportunity to study the coronal properties of a single star which has dissipated its accretion disk and is approaching the ZAMS. This study also allows us to compare the coronal properties of PZ~Tel with those of both younger and older stars, providing a glimpse of the evolution of stellar coronae through the PMS phase. In particular, the CTTSs have disks from which active accretion is still taking place. These stars could have quite different magnetospheric geometries, possibly involving magnetic connections between star and disk \\citep[e.g.][]{Montmerle2002}. In the case of coeval and older stars, it appears that coronal activity of PMS stars without an accreting disk can be explained on the same basis as that of main-sequence stars \\citep*{FlaccomioMicela2003} --- stellar rotation and convection. In this domain, PZ~Tel provides a new window on phenomena such as the chemical fractionation of elements that is seen to occur in coronae over a wide range of activity level. In the solar case, elements with low first ionization potential (FIP) such as Mg, Fe and Si are seen to be enhanced relative to elements with high FIP, such as O, Ne and Ar \\citep[e.g.][]{Feldman1992}. In more active stars, the situation appears somewhat reversed, with elements such as Ne appearing enhanced relative to those with lower FIP \\citep[e.g.][and references therein]{Drake2002}. The case of the active but very young post-T Tauri stars remains unexplored at high spectral resolution. In \\S~\\ref{obs} and \\ref{analysis} we describe the {\\em Chandra} observation and the techniques used in its analysis. Section~\\ref{results} presents the resulting coronal temperature structure and abundances. These are discussed in \\S~\\ref{disc}, in which we also present a comparison of the coronal properties of PZ~Tel with those of AB~Dor and TW~Hya. \\begin{figure*}[t] \\centering \\includegraphics[width=13cm] {figure1.ps} \\caption{Light curve of PZ~Tel obtained from the HEG and MEG spectra (excluding zero-order events) with bin size of 1000\\,s.} \\label{fig:pztel_lightcurve} \\end{figure*} ", "conclusions": "" }, "0403/astro-ph0403221_arXiv.txt": { "abstract": " ", "introduction": "My visits to Japan have occurred always in crucial times in the development of our research. In 1975 I visited the University of Kyoto and gave the lectures which were then co-authored in the Japanese book with Humitaka Sato \\cite{bookhs}. The focus then was on three major topics: a) the basic paradigm for the identification of a black hole I had just established and which had found a very significant application in Cygnus X-1 through the splendid data obtained by Riccardo Giacconi and Minoru Oda \\cite{bookhs}; b) the Cristodoulou-Ruffini \\cite{CR71} mass-energy formula for black holes: \\begin{equation} E_{BH}^2=M^2c^4=\\left(M_{\\rm ir}c^2 + \\frac{Q^2}{\\rho_+}\\right)^2 + \\frac{L^2c^2}{\\rho_+^2}\\, , \\label{em} \\end{equation} where $M_{\\rm ir}$ is the irreducible mass, $\\rho_+ = 2(G/c^2) M_{\\rm ir}$ is the quasi-spheroidal cylindrical coordinate of the horizon evaluated at the equatorial plane and $Q$ and $L$ are respectively the charge and angular momentum of the black hole; this mass-energy formula allows to estimate the maximum energy extractable from a process of gravitational collapse; c) a specific energy extraction process from the black hole by pair creation due to supercritical electric fields, first introduced by Sauter \\cite{S31}, Heinsenberg \\& Euler \\cite{HE35}, Schwinger \\cite{S51}, I developed with T. Damour \\cite{DR75}. In that paper we had also pointed out that such process could be the source of the then newly discovered Gamma-Ray Bursts (GRBs). Our model had a very distinct signature, which differentiates it from all the other models: the characteristic energy of the GRBs should be of the order of $10^{54}$ ergs (see Fig. \\ref{fig1a}--\\ref{fig1b}). \\begin{figure}[t] \\begin{center} \\includegraphics[width=8cm,clip]{foto} \\end{center} \\caption{Princeton 1971.} \\label{fig1a} \\end{figure} \\begin{figure}[t] \\begin{center} \\includegraphics[width=\\hsize,clip]{dia3a} \\end{center} \\caption{The basic components of the Damour \\& Ruffini black hole vacuum polarization.} \\label{fig1b} \\end{figure} The strategy we had followed, both in the case of Cygnus X1 and the GRBs, was not to try to understand the astrophysical aspects of the phenomenon evidencing the black hole formation. On the contrary we had capitalized on the physics of the black hole and on specific properties of the solutions of Einstein Maxwell Equations in order to infer specific signatures to be expected in the astrophysical scenario in order to obtain the observational evidence for a black hole in a realistic astrophysical setting. Indeed the paradigm for the identification of the black hole in Cygnus X1 (Leach \\& Ruffini \\cite{lr73}) was based mainly on three general relativistic considerations: a) the comprehension of the gravitational binding energies around a Kerr black hole, which I found with Wheeler in 1969 \\cite{ll}, clearly pointing to the possibility of having accretion energy as the origin of the observed enormous luminosities in X-ray observed in Cygnus X1, $L=10^4L_\\odot$; b) the uniqueness theorem of black hole (see e.g. Ruffini \\& Wheeler \\cite{rw71}) endowed only of charge mass and angular momentum, clearly pointing to the impossibility of having periodic signals out of a black hole; c) the existence of an absolute maximum mass of a neutron star, again derived out of first principles, from the equation of equilibrium in the Einstein theory of gravity, the principle of causality implying speed of sound not exceeding the speed of light, and the existence of a fiducial density (Rhoades \\& Ruffini \\cite{rr74}). All these points were later summarized in the proceedings of the Varenna School organized by Riccardo Giacconi and myself \\cite{gr75,gr78} and in the Solvay conference \\cite{r74}. Riccardo Giacconi, in his splendid lecture \\cite{nobel} recalls the significance of this theoretical work for the understanding of binary X-ray sources. In the case of GRBs our approach was similar: priority was given to the identification of the energy source of GRBs. That nuclear energy is the energy source of main sequence stars has been credibly proved \\cite{sch}, that accretion and gravitational energy release around neutron stars and black holes was the Energy sources of binary x-ray sources had been demonstrated \\cite{nobel}, we decided to look in the possibility of having a new energy source as powering the GRBs: the extractable energy of a black hole \\cite{CR71,DR75}. The mechanism I had conceived with T. Damour was indeed viable, supported by the very basic and well established physical principles on which it was grounded. My second scientific visit to Japan occurred in occasion of the sixtieth birthday of Humitaka Sato: again I reported \\cite{rukyoto} some new progress in our research: a) The situation with GRBs had dramatically modified by the observations of the Italian-Dutch satellite BeppoSAX (Costa \\cite{ca97}) which gave origin to an unprecedented collaboration between X- and $\\gamma$-ray, optical and radio astronomy. This observational effort had lead to the determination of the distances of GRBs had unequivocally established the cosmological nature of their source: indeed energetics of the order of $10^{54}$ ergs were implied as predicted by our model with Damour \\cite{DR75}. b) I clarified some basic conceptual issues on the energy extraction process from a black hole endowed with electromagnetic structure, introducing the novel concept of ``dyadosphere'' of a black hole, as the region surrounding the black hole horizon where the electron-positron pairs created in the process of vacuum polarization are localized. c) I finally pointed out that the very process of thermalization of such an electron-positron plasma created in the dyadosphere is the main mechanism originating the GRB expansion and the engine of the entire GRB phenomenon \\cite{RSWX99,RSWX00} (see Fig. \\ref{fig2a}--\\ref{fig2b}). \\begin{figure}[t] \\begin{center} \\includegraphics[width=\\hsize,clip]{dia6a} \\end{center} \\caption{The basic parameters of the dyadosphere.} \\label{fig2a} \\end{figure} \\begin{figure}[t] \\begin{center} \\includegraphics[width=8cm,clip]{cl3d} \\end{center} \\caption{The theoretically predicted luminosity and spectral distribution of a short GRB. Details in Ruffini et al. \\cite{RFVX03}.} \\label{fig2b} \\end{figure} In this visit I like to report new results connected with: a) the physics of the dyadosphere and a possible future verification derived from the observation of short GRBs; b) the three fundamental paradigms for the theoretical interpretation of GRBs; c) our understanding of the long bursts and the GRBs afterglows. ", "conclusions": "" }, "0403/astro-ph0403198_arXiv.txt": { "abstract": " ", "introduction": "A number of observations of BHC's do not show consistent correlations of low QPO frequency with disk parameters [1,2]. On the other hand, strong consistent correlations between power law spectral index and low frequency QPO's have been recently observed [3]. In addition, there is mounting observational evidence that the large number of spectral \"states\" formerly developed by to explain the wide variability of BHC's such as GRS1915+105, can be reduced to a few canonical states, i.e. a hard state with spectral power law index $\\Gamma \\sim1.6\\pm 0.1$ , a soft or \"extended\" power-law state characterized by $\\Gamma \\sim 2.7\\pm0.2$, and a thermal state [4]. \\par These data have prompted us to introduce a model, i.e. the Transition Layer (TL) model [5] to explain the correlations observed. The main feature of the TL model is a hot compact region near the BH which serves as the primary region for Compton upscattering of soft disk photons. The TL model shows how the QPO's are related to the size, optical depth, temperature and spectral index and predicts the correlation between index and QPO frequency. ", "conclusions": "" }, "0403/astro-ph0403151_arXiv.txt": { "abstract": "We present Very Large Array observations at 7 mm of the sources IRAS 2A, IRAS 2B, MMS2, MMS3 and SVS 13, in the NGC1333 region. SVS 13 is a young close binary system whose components are separated by 65 AU in projection. Our high angular resolution observations reveal that only one of the components of the SVS 13 system (VLA 4B) is associated with detectable circumstellar dust emission. This result is in contrast with the well known case of L1551~IRS5, a binary system of two protostars separated by 45 AU, where each component is associated with a disk of dust. Both in SVS 13 and in L1551~IRS5 the emission apparently arises from compact accretion disks, smaller than those observed around single stars, but still massive enough to form planetary systems like the solar one. These observational results confirm that the formation of planets can occur in close binary systems, either in one or in both components of the system, depending on the specific angular momentum of the infalling material. ", "introduction": "SVS~13, in the NGC1333 region, was discovered as a 2.2 $\\mu$m source by Strom, Vrba, \\& Strom (1976), and since the source is roughly aligned with the chain of Herbig-Haro objects 7-11 (Herbig 1974; Strom, Grasdalen, \\& Strom 1974), it was assumed to be the exciting source of this classical HH system. Later, Goodrich (1986) detected a faint visible counterpart of SVS~13. However, the star SVS~13 presents a number of peculiar properties. The source exhibited a significant increase of its brightness at optical ($\\sim$3 mag), and IR ($\\sim$1 mag) wavelengths in 1988-1990 (Eisl\\\"offel et al. 1991; Liseau, Lorenzetti, \\& Molinari 1992; Harvey et al. 1998), and since then, the flux has remained almost steady (Aspin \\& Sandell 1994; Khanzadyan et al. 2003). In addition, despite being optically visible, indicating that it is a relatively evolved young object, SVS 13 is a strong millimeter source, known as MMS1 (e.g., Grossman et al. 1987; Looney, Mundy, \\& Welch 2000), and presents other characteristics, such as the presence of an extremely high velocity CO outflow, that suggest it is in a much earlier evolutionary stage (a Class 0/I object; Bachiller et al. 2000). Radio continuum emission from SVS 13 was first reported by Snell \\& Bally (1986). Rodr\\'{\\i}guez, Anglada, \\& Curiel (1997, 1999) mapped SVS 13 with the Very Large Array (VLA) at 3.6 and 6 cm (their source VLA 4). Anglada, Rodr\\'{\\i}guez, \\& Torrelles (2000), through VLA observations at 3.6 cm of higher angular resolution and sensitivity, discovered that SVS~13 is, in fact, a close binary system. The two components of the binary (VLA 4A and VLA 4B) are separated by $0\\rlap.''3$, corresponding to 65 AU in projection (assuming a distance of 220 pc; \\v{C}ernis 1990), and have similar flux densities at 3.6 cm. The water masers associated with SVS~13 appear segregated in position and velocity, supporting the binary hypothesis (Rodr\\'{\\i}guez et al. 2002). Anglada et al. (2000) noted that the optical position for SVS~13 (as measured by Rodr\\'{\\i}guez et al. 1997) is closer to VLA~4A, while the millimeter position (Looney et al. 2000) is closer to VLA~4B. Although the precision of the astrometry available at that time did not allow an unambiguous association, this result led these authors to suggest that the strong millimeter emission reported for SVS~13 could arise from only one of the components of the binary (VLA~4B, the eastern component), while the optical emission would come from the other component (likely from VLA~4A, the western component). In the interpretation proposed by Anglada et al. (2000), only one of the stars (VLA 4B) is surrounded by a dusty envelope or disk, while the other (VLA 4A, the visible star) is not. A similar interpretation has been proposed recently by Loinard et al. (2002) for IRAS 04368+2557 in L1527, on the basis of the morphology of the two obscured sources observed at 7 mm. A confirmation of this interpretation for SVS 13 requires a precise comparison of the positions of the sources observed at different wavelengths, in order to identify the individual contribution of each component of the binary. Since the angular separation between VLA 4A and VLA 4B is only $0\\rlap.''3$, absolute astrometry down to $<0\\rlap.''1$ is required in order to obtain an accurate enough registration of the positions. Although the accuracy of the absolute astrometry of the optical observations ($\\pm0\\rlap.''3$) is difficult to improve due to the lack of a large enough number of suitable reference stars in the field, it is possible to improve the accuracy of the registration between the centimeter and millimeter positions by using the same instrument and calibration procedures in both wavelength ranges. In this Letter, we present VLA observations at 7 mm of the region near SVS 13, carried out in the D and B configurations, using the same phase calibrator and procedures as in the previous VLA observations at 3.6 cm. The B configuration observations at 7 mm provide an angular resolution of $\\sim0\\rlap.''2$, similar to that of the A configuration at 3.6 cm, and an expected accuracy in the registration between both images down to $<0\\rlap.''05$, allowing a precise comparison of the emission at both wavelengths, necessary to test the single disk hypothesis for the SVS 13 binary. These observations also provide 7 mm data on other sources in NGC1333: VLA 2 (MMS3), VLA 7 (IRAS 2A), VLA 10 (IRAS 2B), and VLA 17 (MMS2). ", "conclusions": "We detect at 7 mm the sources VLA 2 (MMS3), VLA 7 (IRAS 2A), VLA 10 (IRAS 2B), and VLA 17 (MMS2=SVS 13B) (see Table 1). These sources were previously observed at 3.6 and 6 cm by Rodr\\'\\i guez et al. (1999), where a discussion on their properties and counterparts can be found. VLA 7, VLA 10, and VLA 2 were further observed at 3.6 cm with higher angular resolution by Reipurth et al. (2002). Interferometric observations at 3 mm of VLA 7 and VLA 10 have been recently reported by J{\\o}rgensen et al. (2004). We also detect SVS 13 at 7 mm, as an unresolved source in the D configuration, and resolving it in its two components (VLA 4A and VLA 4 B) in the B configuration observation (see Table 1). In Figure 1 we compare the 3.6 cm map observed with the A configuration (Anglada et al. 2000) with the 7 mm map observed with the B configuration (this paper). Both maps were obtained with a similar angular resolution of $\\sim 0\\rlap.''2$. The positions of the sources in the two maps are in agreement within $0\\rlap.''02$, eliminating, thus, any ambiguity in their identification at different wavelengths. As can be seen in the figure, at 3.6 cm both sources present a similar flux density, while at 7 mm VLA 4B is much stronger than VLA 4A, suggesting that VLA 4B is the dominant source in the millimeter wavelength range, with a negligible contribution from VLA 4A. In order to quantitatively confirm this hypothesis it should be verified that the increase of flux density of VLA 4B over VLA 4A at 7 mm is caused by dust emission, and not by free-free emission with a steep spectral index. Usually, the emission at wavelengths longer than a few cm is dominated by the free-free emission from ionized gas, while the emission at wavelengths shorter than a few mm is dominated by thermal emission from dust. Since the wavelength of 7 mm falls in between the centimeter and millimeter regimes, it should be checked, for each particular object, which is the nature of the dominant emission at this wavelength. To do that, we have plotted in Figure 2 the highest angular resolution data available on SVS 13 in the centimeter and millimeter ranges. The flux densities of VLA 4B at 3.6 cm and 1.3 cm give a spectral index of $\\alpha=1.36\\pm0.26$, which is typical of a thermal ionized jet or a partially optically thick ionized region. An extrapolation at 7 mm of this free-free emission yields a flux density of $\\sim$1 mJy, much smaller than the observed value. On the other hand, a fit to the observed data from 3.4 mm to 1.3 mm (that are supposed to trace dust emission) gives a spectral index $\\alpha=2.55\\pm0.05$ in the millimeter range, resulting in an extrapolated flux density at 7 mm in good agreement with the value observed for VLA 4B (see Table 1 and Fig. 2). Then, we conclude that free-free emission cannot account for the observed flux density of VLA 4B at 7 mm, and that a contribution of a different nature, namely, thermal dust emission, likely from a circumstellar disk surrounding this object, is required. In the case of VLA 4A, the overall 3.6 cm, 1.3 cm, and 7 mm data points can be fitted together with a single spectral index $\\alpha=1.25\\pm0.20$, indicating that the observed flux density at 7 mm can be explained as free-free emission (see Fig. 2). However, we cannot discard a small contribution from dust emission at 7 mm, since an extrapolation of the 3.6 and 1.3 cm data alone gives a free-free contribution at 7 mm slightly below the observed flux density of VLA 4A. The difference ($\\la$0.8 mJy) could correspond to a contribution from dust, being it about five times smaller than in the case of VLA 4B. In summary, our results indicate that of the two components of the SVS 13 binary system, the source VLA 4B is associated with a much larger amount of dust than the source VLA 4A. This strongly supports the proposal by Anglada et al. (2000), who suggested that the millimeter source MMS1 is the counterpart of the centimeter source VLA 4B. On the basis of the available optical astrometry, these authors also proposed that VLA 4A is the counterpart of the visible star SVS 13. VLA 4B appears as a compact source in our 7 mm map obtained with the B configuration with an angular resolution of $\\sim0\\rlap.''2$ (Fig. 1b), suggesting that the dust emission traced by this source originates in a compact circumstellar structure, likely a disk, with radius $\\la 30$ AU. The size of the disk is smaller than that of the typical accretion disks observed around single T Tauri stars, with radii of 100-150 AU (e.g., Dutrey et al. 1996; Wilner et al. 2000), and is more similar to that of the compact disks observed in the L1551~IRS5 binary system (10 AU; Rodr\\'\\i guez et al. 1998), suggesting that they are truncated by the tidal effects of the companion star. The 7 mm flux density of VLA 4B, together with the millimeter data shown in Figure 2, can be fitted by a simple disk model. We assume a geometrically thin, vertically isothermal disk, characterized by power-law radial dependences of temperature and surface density, and we adopt the opacity law given by D'Alessio, Calvet, \\& Hartmann (2001). The data are fitted with a distribution of dust grains with a maximum size of 1 mm, a temperature distribution $T(r)=470~(r/\\rm AU)^{-0.5}$~K, a surface density $\\Sigma(r)=2800~(r/\\rm AU)^{-1}$~g~cm$^{-2}$, implying a disk mass of 0.06 $M_\\odot$, for a radius of 30 AU. Thus, it seems plausible that the emission of VLA 4B is tracing a protoplanetary disk, since the mass obtained exceeds the minimum mass required to form a planetary system like the solar one ($\\sim 0.01~M_\\odot$). Since any dust emission associated with VLA 4A is at least five times weaker than that of VLA 4B, if a disk was associated with VLA 4A, we expect its mass to be at least five times smaller than that of the disk associated with VLA 4B. We note that the flux density observed in the D configuration is larger than the total flux observed in the B configuration, indicating an additional contribution from an extended component, likely the infalling envelope. A fit to the overall spectral energy distribution, taking into account simultaneously the contributions of different components (disks+envelope), similarly to what has been done for L1551 IRS5 (Osorio et al. 2003), would be useful to better constrain the properties of each component. Thus, the observational results obtained for SVS 13 imply that it is feasible that the development of a protoplanetary disk occurs preferentially in only one of the components of a young close binary system, with the disk absent or much less significant in the other component. In this respect, the case of the SVS 13 binary system appears to be opposite to the L1551~IRS5 case, where both components of the binary system are associated with circumstellar disks of dust of comparable characteristics (Rodr\\'\\i guez et al. 1998). These results are in agreement with theoretical simulations (e.g., Bate \\& Bonnell 1997) that show that, depending on the mass ratio of the components and the specific angular momentum of the system, the development of circumstellar disks can occur either around a single component or around both components of the binary. According to Bate \\& Bonnell (1997), a circumstellar disk forms around one of the components of the binary only if the specific angular momentum of the infalling gas is greater than the specific orbital angular momentum of that component about the center of mass of the binary. Thus, if a binary system grows to its final mass mainly via the accretion of material with low specific angular momentum, the primary may have a large circumstellar disk, while the secondary is essentially naked. The systems formed by this method are expected to be binaries with separations of the order of $\\sim 100$ AU, and they should not be developing a significant circumbinary disk. For infall with high angular momentum, both components can develop a circumstellar disk and even a circumbinary disk can be formed. Under this simplified scheme, the low angular momentum scenario, with a disk in only one star, appears to correspond to the case of the SVS 13 binary, where the two components are separated by 65 AU in projection, and where VLA 4A would be the secondary, whereas VLA 4B (the dominant source at millimeter wavelengths) would be the primary viewed through its circumstellar disk. On the other hand, L1551~IRS5 apparently corresponds to a case of accretion of material with higher angular momentum, with both components associated with a circumstellar disk (Rodr\\'\\i guez et al. 1998) and probably surrounded by a circumbinary disk (Osorio et al. 2003). In order to confirm these suggestions, it would be interesting to obtain the orbital parameters of these systems and to identify which component is the primary through accurate measurements of absolute proper motions. Relative proper motions between the two components of the L1551~IRS5 system have been obtained recently (Rodr\\'\\i guez et al. 2003), and accurate absolute proper motions have been measured for the T Tau Sa/Sb system (Loinard, Rodr\\'\\i guez \\& Rodr\\'\\i guez 2003). At present, no proper motions are available for the SVS 13 system, but these promising results obtained for other sources suggest that this goal could be attainable in a relatively near future, providing a complete test for the properties of the SVS 13 binary system." }, "0403/astro-ph0403367_arXiv.txt": { "abstract": "A semianalytic method to estimate the angular resolution of tracks, that have been reconstructed by a likelihood approach, is presented. The optimal choice of coordinate systems and resolution parameters, as well as tests of the method are discussed based on an application for a neutrino telescope. ", "introduction": "This paper describes a statistical procedure to extract resolution estimates on a track by track basis. The method was developed for the AMANDA Neutrino Telescope at the South Pole~\\cite{tillpub:Andres:2001ty}, which uses a 3-dimensional grid of photosensors imbedded in highly transparent ice to provide spatial and time resolution of Cherenkov photons, that e.g. arise from long muon tracks. The knowledge of track resolutions is of particular importance in the search for localized sources, such as distant galaxies. The resolution information can in addition be used to suppress mis-reconstructed tracks, that typically are less well defined. The method is not limited to muon reconstruction in neutrino telescopes and can be applied to any experiment in which tracks have been reconstructed with a likelihood approach. ", "conclusions": "\\label{section:tests} The above algorithms have been tested within the framework of the AMANDA neutrino telescope. The estimates perform in a stable and reliable way and produce sensible results. To ensure the correctness of the estimations, several testing procedures both in data and in Monte Carlo have been carried out: \\begin{enumerate} \\item For $\\sigma_{\\theta}$ and $\\sigma_{\\phi}$ its {\\em pull\\/} has been studied. It is defined as the ratio of the difference of true and reconstructed direction over the resolution estimator. If the estimation is good, the pull should be Gaussian distributed, centered at zero and with unit width. This investigation can only be done in a Monte Carlo simulation. \\item For the spatial angle resolution the pull is not a sensible quantity, because angles in space can only be positive. Hence one can study ensembles of events with the same $\\sigma_{\\mathrm{a}}$ or $\\sigma_{\\mathrm{a}}^{\\epsilon}$, respectively. The median of the true deviation can then be compared to the corresponding $\\sigma_{\\mathrm{a}}$. Again this can only be done in the simulation. \\item In data the true directions are not known. Thus another approach is being followed. Each event is split into two subevents by assigning every second hit to subevent 1 and the remaining hits to subevent 2. Each subevent undergoes reconstruction and is subjected to the resolution estimation algorithm. That leads to directions\\footnote{Note that the reconstructed directions are expressed in standard detector coordinates, whereas the error estimates are obtained in their respective rotated systems. That is due to technical reasons only.} $\\vartheta^i$, $\\varphi^i$ and errors $\\sigma_{\\theta}^i$ und $\\sigma_{\\phi}^i$ with $i\\in\\{1,2\\}$. The pull \\begin{equation} \\mathrm{P}_{\\vartheta} = \\frac{\\mathrm{D}_{\\vartheta}}{\\sigma_{\\theta}^{\\mathrm{D}}} = \\frac {\\vartheta^1 - \\vartheta^2} {\\sqrt{(\\sigma_{\\theta}^1)^2 + (\\sigma_{\\theta}^2)^2}} \\end{equation} should also be a Gaussian distribution centered at zero with unit width. For $\\mathrm{P}_{\\varphi}$ the difference in azimuth angles must be multiplied by $\\sin\\vartheta$ to make up for the differences in the rotated and not rotated coordinate systems: \\begin{equation} \\mathrm{P}_{\\varphi} = \\frac{\\mathrm{D}_{\\varphi}}{\\sigma_{\\phi}^{\\mathrm{D}}} = \\frac {(\\varphi^1 - \\varphi^2)\\cdot\\sin\\vartheta} {\\sqrt{(\\sigma_{\\phi}^1)^2 + (\\sigma_{\\phi}^2)^2}} \\qquad . \\end{equation} Of course this last check can also be done in the simulation. \\end{enumerate} In all tests the results were satisfactory. Further details about the tests and their results can be found in \\cite{phd:neunhoeffer_eng:2003}.% \\subsection{Concluding remarks} \\begin{enumerate} \\item Often one applies quality cuts to the data set, which rely on additional information that is not reflected in the construction of the likelihood function itself. Hence this information is also not used in obtaining the resolution parameters. In this case one expects the selected tracks to be on average closer to the true direction than from the likelihood analysis alone. Consequently one observes a pull with a width smaller than 1. \\item The resolution parameters - such as $\\sigma_{\\mathrm{a}}$ - can efficiently be used as cut variables, as misreconstructed events on average have a worse resolution than correctly reconstructed ones. \\item The resolution estimation on an event-per-event basis can be used as additional input in point source searches. The information is hard to be included in the commonly used binned search algorithms \\cite{tillpub:ahrens:2003xy3}. An appropriate procedure based on maximum likelihood methods that can integrate the new information in a natural way will be presented in a forthcoming publication. \\end{enumerate} \\begin{ack} I wish to thank the AMANDA collaboration for their support in obtaining the method described herein. I thank Lutz K{\\\"o}pke and Alexander Holfter for many an illuminative and productive discussion. I would also like to thank the German Research Foundation (DFG) and the German Ministry of Research and Education (BMBF) for financial support of the AMANDA project. \\end{ack}" }, "0403/astro-ph0403684_arXiv.txt": { "abstract": "We present an analysis of candidate members of the $\\eta$~Cha and MBM~12A young associations. For an area of 0.7~deg$^2$ toward $\\eta$~Cha, we have performed a search for members of the association by combining $JHK_s$ photometry from 2MASS and $i$ photometry from DENIS with followup optical spectroscopy at Magellan Observatory. We report the discovery of three new members with spectral types of M5.25-M5.75, corresponding to masses of 0.13-0.08~$M_{\\odot}$ by theoretical evolutionary models. Two and three of these members were found independently by Lyo and coworkers and Song and coworkers, respectively. Meanwhile, no brown dwarfs were detected in $\\eta$~Cha down to the completeness limit of 0.015~$M_{\\odot}$. For MBM~12A, we have obtained spectra of three of the remaining candidate members that lacked spectroscopy at the end of the survey by Luhman, all of which are found to be field M dwarfs. Ogura and coworkers have recently presented four ``probable\" members of MBM~12A. However, two of these objects were previously classified as field dwarfs by the spectroscopy of Luhman. In this work, we find that the other two objects are field dwarfs as well. ", "introduction": "The environs of the B8 star $\\eta$ Cha and the dark cloud MBM~12 have been revealed as sites of small associations of newly formed stars. Through deep {\\it ROSAT} observations of 0.35~deg$^2$ toward $\\eta$ Cha, \\citet{mam99} discovered X-ray emission from 12 sources, including $\\eta$ Cha itself. From {\\it Hipparcos} parallactic distances and proper motions, the two earliest stellar counterparts and an early-type star without X-ray emission were found to be comoving at a distance of $\\sim100$~pc. \\citet{mam99} confirmed the youth of the remaining 10 late-type stars through measurements of H$\\alpha$ emission and Li absorption, which indicated their probable membership in an association with the early-type stars. An age of 4-10 Myr has been inferred for these stars through comparisons of their positions on the Hertzsprung-Russell (H-R) diagram to theoretical isochrones \\citep{mam99,law01}. Recently, five additional low-mass members of the association (0.08-0.3~$M_{\\odot}$) have been identified through spectroscopy of candidates appearing in color-magnitude diagrams generated from photometry at $V$ and $I$ \\citep{law02,lyo04} and from data in the USNO-B1 and Two-Micron All-Sky Survey (2MASS) catalogs \\citep{sz04}. Although not precisely determined, the completeness limits of these latest surveys appeared to be near 0.1~$M_{\\odot}$. Surveys for sources of H$\\alpha$ and X-ray emission toward cloud 12 from \\citet{mbm85} (MBM~12) have resulted in the discovery of seven T Tauri stars (e.g., \\citet{hb88}). MBM~12 was long believed to be the nearest molecular cloud at a distance of 50-100 pc \\citep{hbm86,hea00}, making the associated group of young stars (MBM~12A) a particularly attractive site for studies of star formation. However, during a search for new members, \\citet{luh01} noticed that the members of MBM~12A exhibited anomalously old ages ($\\sim100$~Myr) when placed on the H-R diagram with the published distances. Through new distance estimates in that work and subsequent studies \\citep{str02,and02}, the distance of MBM~12 cloud is now firmly established near a value of $\\sim300$~pc. Using this distance, theoretical isochrones produce an age of $\\sim2$~Myr, which is consistent with the evolutionary state of the association implied by various signatures of newly formed stars \\citep{luh01}. The magnitude-limited survey by \\citet{luh01} achieved a completeness limit of 0.03~$M_{\\odot}$ and uncovered five new members (0.1-0.4~$M_{\\odot}$), bringing the total known membership to a dozen sources at a resolution of $\\sim1\\arcsec$. \\citet{ogu03} have since identified four ``probable\" members of MBM~12A via H$\\alpha$ emission detected with slitless grism spectroscopy. In this paper, we present spectroscopy of candidate members of the $\\eta$~Cha and MBM~12 associations. For $\\eta$~Cha, we select candidate members of the association by combining $i$ photometry from the Deep Near-Infrared Survey of the Southern Sky (DENIS) and $JHK_s$ data from the Two-Micron All-Sky Survey (2MASS) (\\S~\\ref{sec:ident1}), measure their spectral types and determine their status as field stars or association members (\\S~\\ref{sec:class1}), evaluate the completeness of the survey (\\S~\\ref{sec:complete}), and place the known members of $\\eta$~Cha on the H-R diagram (\\S~\\ref{sec:hr}). For MBM~12A, we select candidate members from \\citet{luh01} and \\citet{ogu03} for spectroscopy (\\S~\\ref{sec:ident2}) and use these data to measure spectral types and assess membership (\\S~\\ref{sec:class2}). ", "conclusions": "We have performed spectroscopic studies of candidate members of the $\\eta$~Cha and MBM~12A young associations, the conclusions for which are summarized as follows: For $\\eta$~Cha, we have used $JHK_s$ photometry from 2MASS and $i$ photometry from DENIS to construct color-color and color-magnitude diagrams for an area of 0.7~deg$^2$. Through spectroscopy of the candidate members appearing in these data, we have discovered three new members of the association with spectral types of M5.25-M5.75, corresponding to masses of 0.13-0.08~$M_{\\odot}$ according to evolutionary models. These sources were independently found in recent work by \\citet{sz04} and \\citet{lyo04}. No brown dwarfs were detected in $\\eta$~Cha down to the completeness limit of 0.015~$M_{\\odot}$ for our survey, which is roughly consistent with the yield of $\\sim2$ substellar members expected if the relative numbers of stars and brown dwarfs in $\\eta$~Cha are similar to those of the Taurus and IC~348 star-forming regions. On the other hand, the fact that the three least massive members of the association are the outermost members may indicate that the substellar members are preferentially located at large distances from the center of the association and outside of our survey field. For the 18 known members of $\\eta$~Cha, we have estimated bolometric luminosities and effective temperatures and placed the members on the H-R diagram, from which we infer an age of $6^{+2}_{-1}$~Myr for the association with the evolutionary models of \\citet{bar98}. For MBM~12A, we have presented spectra of three of the remaining candidate members that lacked spectroscopy at the end of the survey by \\citet{luh01} and of the four ``probable\" members from \\citet{ogu03}. We classify all of these sources as field dwarfs based on the absence of Li absorption in their spectra." }, "0403/astro-ph0403401_arXiv.txt": { "abstract": "{ We present a statistical analysis of the {\\it Chandra} observation of the source field around the 3C 295 galaxy cluster (z=0.46) aimed at the search for clustering of X-ray sources. We applied three different methods of analysis, all suggesting a strong clustering in the field on scales of a few arcmin. In particular 1) the logN-logS computed separately for the four ACIS-I chips reveals that there is a significant ($3.2\\; \\sigma$ in the $0.5-2$ keV, $3.3\\; \\sigma$ in the $2-10$ keV and $4.0\\; \\sigma$ in the $0.5-10$ keV band) excess of sources to the North-North East and a void to the South of the central cluster. 2) the two point, two-dimensional Kolmogorov-Smirnov (KS) test, shows the probability that the sources are uniformly distributed is only a few percent. 3) a strong spatial correlation emerges from the study of the angular correlation function of the field: the angular correlation function (ACF) shows a clear signal on scales of $0.5\\div 5$ arcmin, correlation angle in the $0.5-7$ keV band $\\theta_0=8.5^{+6.5}_{-4.5}$, $90$\\% confidence limit (assuming a power law ACF with slope $\\gamma=1.8$). This correlation angle is $2$ times higher than that of a sample of $8$ ACIS-I field at the $2.5 \\; \\sigma$ confidence level. The above scales translate to 0.2$\\div$2 Mpc at the cluster redshift, higher than the typical cluster core radius, and more similar to the size of a ``filament'' of the large scale structure. ", "introduction": "N-body and hydrodynamical simulations show that clusters of galaxies lie at the nexus of several filaments of galaxies (see e.g. Peacock 1999, Dav\\`e et al. 2001 and references therein). Such filaments map out the ``cosmic web'' of voids and filaments of the large scale structure (LSS) of the Universe. According to the same simulations these filaments contain a large fraction ($30-40\\%$) of the baryons in the Universe at z$<1$ (the remainder ending up in the hot gas in clusters of galaxies on one side, and in stars and cold gas clouds on the other side). Despite its larger total mass, observations of the intergalactic matter in filaments have yielded so far only limited information, mostly due to its low density (most of the baryons in this phase should be at densities only 10-100 times higher than the average density in the Universe). The most direct ways to detect a filament at low redshift is through its soft X-ray diffuse emission (see e.g. Zappacosta et al. 2002, Soltan, Freyberg \\& Hasinger 2002), or through soft X-ray and UV absorption line studies (see e.g. Fiore et al. 2000 and references therein, Nicastro et al. 2002, Nicastro et al. 2003, Mathur et al. 2003). Both methods require very difficult observations, at the limit of the present generation of X-ray and UV facilities. Alternatively, filaments could be mapped out by galaxies (Daddi et al. 2001, Giavalisco \\& Dickinson 2001) and by the much more luminous Active Galactic Nuclei (AGNs), assuming that AGNs trace galaxies. Since rich clusters of galaxies are good indicators of regions of sky where filaments converge, numerous AGN searches around clusters of galaxies have been performed in the past. Several of these studies suggest overdensities of AGNs around distant clusters of galaxies (Molnar et al. 2002 for the cluster Abell 1995; Best et al. 2002 for MS1054-03; Martini et al. 2002 for Abell 2104; Pentericci et al. 2002 for the protocluster at $z \\sim 2.16$ around the radio galaxy MRC 1138-206; see also Almaini et al. 2003 for the ELAIS North field). Many of these studies have been performed in X-rays, since extragalactic X-ray sources, which are mostly AGNs, have a space density $\\sim 10$ times higher than optically selected AGNs (see Yang et al. 2003), and therefore provide denser tracing of LSS. One of the first studies of the X-ray source population in cluster fields was performed by Cappi et al. (2001) who studied the {\\it Chandra} $8\\times 8$ arcmin ACIS fields around the distant cluster of galaxies RX J003033.2+261819 ($z=0.5$) and 3C 295 ($z=0.46$). Cappi et al. (2001) reported the tentative detection of an overdensity of faint X-ray sources in a region of a few arcmin around both clusters, with respect to the average X-ray source density at the same flux limit. However, the observations were too short ($\\sim 30$ ks and $\\sim 18$ ks, respectively), and the source samples consequently too small, to derive any more detailed conclusion. In addition to counting sources in selected sky areas, the clustering of X-ray sources can be studied using other observational and statistical tools. Gilli et al. (2003) report narrow spikes in the redshift distribution of the sources in the {\\it Chandra} Deep Field South (CDFS), indicating strong clustering of sources in these narrow redshift ranges. If the source samples are sufficiently large (of the order of 100 sources or more) one derives more detailed and quantitative information on the source clustering by studying the angular correlation function (ACF) of the sources in the field. Vikhlinin \\& Forman (1995) were among the first to study the ACF in the X-ray band; they evaluated an average ACF from a large set of deep ROSAT observations, covering in total 40 deg$^2$ of sky. They found positive correlation on scales from a fraction of arcmin to tens of arcmin. However, their best fit ``correlation angle'' $\\theta_0 \\sim 10$ arcsec, is smaller than the ROSAT PSPC Point Spread Function ($\\sim 25$ arcsec FWHM on-axis); as noted by the authors, this leads to an ``amplification bias'' and their measured ACF is consequently somewhat overestimated. This effect can be greatly mitigated by using a telescope like {\\it Chandra}, whose on-axis PSF is only 0.5 arcsec FWHM, 50 times better than the PSPC on-axis PSF. So far a correlation analysis of {\\it Chandra} X-ray sources have been published only by Giacconi et al. (2000), who used the first 100 ks of observation of the CDFS, and by Yang et al. (2003) who analyzed a mosaic of 9 moderately deep (30ks) {\\it Chandra} pointings of the Lockman Hole area. Yang et al. could study the source angular correlation on scales of several tens of arcmin, being limited on smaller scales by the small number of detected sources per unit area. Our main goal in this paper is to push this kind of analysis toward a) smaller scales, of order of a few arcmin; and b) targeting a field where the likelihood of observing LSS is high, i.e. a field around a cluster of galaxies. For these reasons we observed again with {\\it Chandra} for about 100ks the 3C295 field, where Cappi et al. (2001) found tentative evidence of an overdensity of faint X-ray sources in a region of a few arcmin around the central cluster. The paper is organized as follows: Section 2 presents the observations and data reduction; Section 3 presents the results of our analysis; Section 4 discusses such results and draws our conclusions. ", "conclusions": "A $92$ ks {\\it Chandra} observation of the source field around the $z =0.46 $ 3C 295 cluster shows an excess of sources visible in the NE corner of the {\\it Chandra} observation (fig. 1). This is clear in the density contour plot of fig. 11. This figure shows that the denser region of the putative filament has a sources density more than $5$ times higher than the average source density of the field ($\\sim 0.5$ sources per arcmin$^2$. A chip by chip logN-logS analysis demonstrates that the NE chip has a $4.0 \\sigma$ excess of sources over the SW chip in the total ($0.5-7$ keV) {\\it Chanda} band. [In the standard soft ($0.5-2$ keV) band the excess is $3.2 \\sigma$, and in the standard hard ($2 - 7$ keV) band the excess is $3.3 \\sigma$.] This result confirms the basic result of Cappi et al. (2001) and extends it to deeper fluxes, larger field-of-view and the $2-10$ keV band. \\begin{figure} \\centering \\includegraphics[angle=0,width=9cm]{cont3.ps}% \\caption{The density profile of the 3C 295 field, computed for the whole $0.5 - 7$ keV band. The linear smoothing factor is $1.5$ arcmin, and the four contour levels indicate source densities of $1.3$, $1.8$, $2.2$ and $2.7$ sources per arcmin$^2$. The clustering of sources in the NE corner is clearly visible; the bright central spot is the 3C 295 cluster. } \\label{spe1} \\end{figure} The asymmetric distribution of the 3C 295 field is confirmed by two analyses: (1) the two dimensional Kolmogorov-Smirnov test, which show the probability that the sources are uniformly distributed is $0.632 \\%$ in the total {\\it Chandra} band [ $3.09\\,\\%$ for the soft band and $3.84\\;\\%$ for the hard band). (2) the two point angular correlation function (ACF) strongly indicate positive correlation, for scales of $0.5-5$ arcmins. This strong correlation has not been found in a sample of eight ACIS-I fields with a similar exposure time, for which the correlation angle $\\theta_0$ is smaller by a factor of $\\gs2$ than that of 3C 295 at the $2.5\\;\\sigma$ confidence level, suggesting that the 3C 295 overdensity of sources in the NE chip is truly peculiar. An intriguing explanation, to be confirmed when the redshifts of the sources are measured, is that the overdensity of sources is actually related to a cosmic filament of the LSS. Cappi et al. (2001) discussed four possible causes for this `surplus' of sources: (1) gravitationally lensed very faint sources; (2) rapid evolution of cluster AGN or starburst galaxies; (3) cosmic variance of background sources; (4) LSS associated with the clusters. Since the surplus sources are not symmetrically place around the cluster, our results rule out an enhanced AGN population in the 3C~295 cluster itself, and lensing by the cluster potential. Since N-body and hydrodynamical simulations and galaxy surveys (e.g 2dF and Sloan) lead us to expect that clusters of galaxies lie at the nexus of several filament, we believe that this excess is likely to represent a filament of the LSS of the Universe converging onto the 3C 295 cluster. The redshifts of the sources making up the excess are not yet known. However, if we assume that the sources are associated with the 3C 295 cluster, then we can use the cluster redshift to estimate some of their properties. Adopting a concordance cosmology ($H_0 =65$ km/s Mpc, $\\Omega_M = 0.3$ and $\\Omega_{\\Lambda} = 0.7$, Spergel et al. 2003) we obtain luminosities in the range $7.5\\times 10^{41} \\div 1.1 \\times 10^{44}$ ergs s$^{-1}$ (median $= 3.1 \\times 10^{42}$ ergs s$^{-1}$) for the total {\\it Chandra} band. [$2.8\\times 10^{41} \\div 3.8\\times10^{43}$ ergs s$^{-1}$ (median $= 1.1 \\times 10^{42}$ ergs s$^{-1}$) for the soft band; $1.5\\times 10^{42} \\div 5.0 \\times 10^{43}$ ergs s$^{-1}$ (median $= 4.5 \\times 10^{42}$ ergs s$^{-1}$) for the hard band.] These are moderate, Seyfert galaxy, luminosities, extending down to the range of luminous starburst galaxies at the faint end of the soft band. An association with 3C 295 also defines a spatial scale of $2$ Mpc (5 arcmin) for the adopted cosmology. For an, admittedly unlikely, spherical region of volume of $\\sim 4 $ Mpc$^{3}$ containing the excess sources, the implied space density of the excess sources (i.e. subtracting the CDF-S defined logN-logS contribution) is $0.9$ Mpc$^{-3}$ ($0.5-2$ keV) and $0.8$ Mpc$^{-3}$ ($2-10$ keV). This is well above the normal maximum AGN space density at $z=0.5$ ($\\sim 10^{-4}$ Mpc$^{-3}$, for luminosities down to $10^{42}$ ergs s$^{-1}$ in the 0.5-2keV band, Hasinger 2003, and $10^{43}$ ergs s$^{-1}$ in the 2-10 keV band Fiore et al. 2003). However, since the X-ray source luminosities are as low as $\\sim 2 \\times 10^{41}$ ergs s$^{-1}$ (at the redshift of 3C 295), the little studied lower end of the XLF is being probed, and contributions from starburst and even normal galaxies can be important. In fact, the integral of the field galaxy luminosity function at z=0.5 (e.g. Poli et al. 2001) gives $\\sim 0.13$ galaxies Mpc$^{-3}$ for $M_B < -17$ (a resonable faint end optical luminosity, corresponding to our lower X-ray luminosities). If we assume that roughly one tenth of the galaxies are active X-ray sources of $L_X > 3 \\times 10^{41}$ ergs s$^{-1}$, then we would expect $\\sim0.013$ X-ray sources Mpc$^{-3}$. Since we count $\\sim0.9$ sources Mpc$^{-3}$, this implies a galaxy overdensity of $\\approx70$, with of course a large (factor of 2-4) positive and negative uncertainty, because of the uncertainties in our space densities and assumptions. Still, this is intriguingly close to the expected galaxy overdensity of filaments $\\sim 10 \\div 10^2$, and much smaller than the overdensities of clusters of galaxies ($\\sim 10^3 \\div 10^4$), (the density contrast at the virial radius being $\\sim200$). Furthermore, in a filamentary structure the implied space density will depend strongly on the orientation of the filament to our line of sight. We stress again that these arguments hold only under the assumption that the redshift of the excess sources is the same of that of the 3C 295 cluster. In order to determine if this excess is associated or not to the 3C 295 cluster, we need to know the redshift of the sources via optical identifications. For this reason, in order to study this candidate filament, we are pursuing new {\\it Chandra} and optical observations to map out the 3C~295 region and delimiting the filament properties up to scales of $\\sim 24$ arcmin (i.e. $\\sim 6$ Mpc) from the 3C 295 cluster. These studies are important because they may open-up a new way to map high-density peaks of LSS at high redshifts with high efficiency." }, "0403/astro-ph0403637_arXiv.txt": { "abstract": "{$uvby$(--$\\beta$) photometry has been obtained for an additional 411 very metal-poor stars selected from the HK survey, and used to derive basic parameters such as interstellar reddenings, metallicities, photometric classifications, distances, and relative ages. Interstellar reddenings adopted from the Schlegel et al.~(1998) maps agree well with those from the intrinsic-color calibration of Schuster \\& Nissen (1989). [Fe/H] values are obtained from the CaII K line index of the HK survey combined with the $uvby$ and $UBV$ photometry. The $c_{\\rm 0},(b-y)_{\\rm 0}$ diagram is seen to be very useful for classifying these very metal-poor field stars into categories similar to those derived from globular cluster color-magnitude diagrams; it is found that the HK survey has detected metal-poor candidates extending from the red-giant to the blue-horizontal branch, and from the horizontal branch to subluminous stars. Distances derived from $UBV$ photometry agree reasonably well with those from $uvby$, considering the paucity of good calibrating stars and the extrapolations required for the most metal-poor stars. These very metal-poor stars are compared to M92 in the $c_{\\rm 0},(b-y)_{\\rm 0}$ diagram, and evidence is seen for field stars 1--3 Gyrs younger than this globular cluster; uncertainties in the [Fe/H] scale for M92 would only tend to increase this age difference, and significant reddening uncertainties for M92 are unlikely but might decrease this difference. The significance of these younger very metal-poor stars is discussed in the context of Galactic evolution, mentioning such possibilities as hierarchical star-formation/mass-infall of very metal-poor material and/or accretion events whereby this material has been acquired from other (dwarf) galaxies with different formation and chemical-enrichment histories. ", "introduction": "Over the past two decades, our collective knowledge of the nature of the thick disk and halo of the Galaxy has expanded enormously, due primarily to the impact of several ongoing large-scale survey efforts carried out to detect and analyze metal-poor stars. These include the HK survey of Beers and collaborators (Beers, Preston, \\& Shectman 1992; Beers 1999) and the Hamburg/ESO stellar survey of Christlieb \\& collaborators (Christlieb 2003), both of which select stars with objective-prism techniques, and hence introduce no kinematic bias into their samples. Such biases are present (and must be corrected for) in proper-motion selected survey samples, such as the exhaustive previous studies of, e.g., Ryan \\& Norris (1991) and Carney et al.~(1996). The prism-survey selected samples are hence well suited for studies of the kinematics and dynamics of the old stellar populations of the Milky Way, in particular because of the burgeoning databases of proper motion information that are presently being assembled from a variety of sources (e.g., UCAC2: Zacharias 2002; SPM: Girard et al.~2003). In order to make optimal use of the proper motions for kinematic analyses, accurate stellar classifications, and photometrically determined distances, are crucial. The $uvby$--$\\beta$ photometric system is particularly suited for the study of very-metal-poor (hereafter, VMP) F- and G-type stars, as has already been pointed out in Paper VIII by Schuster et al.~(1996; hereafter S96). Briefly, intrinsic-color calibrations, $(b-y)_{\\rm 0}$--$\\beta$, exist which allow accurate and precise, $\\pm 0\\fm01$, measures of interstellar reddening excesses, $E(b-y)$, for individual field stars; such a calibration has been given by Schuster \\& Nissen (1989). Photometric absolute magnitudes and distances can be calibrated and used effectively, as shown in the papers by Olsen (1984) and Nissen \\& Schuster (1991). This photometric system has the great advantage that it permits us to obtain accurate stellar distances even for evolving main-sequence and subgiant stars due to the gravity sensitivity of the $c_{\\rm 0}$ index. Also, importantly, theoretical isochrones in the $M_{\\rm bol}$, $T_{\\rm eff}$ diagram can be transformed to the $M_{\\rm V}$, $(b-y)_{\\rm 0}$ or $c_{\\rm 0}$, $(b-y)_{\\rm 0}$ diagrams for the estimation of relative and/or absolute ages of evolving field stars which are near their respective turn-offs, and in several of the previous papers of this series the isochrones of VandenBerg et al.~have been used for such purposes, to study the Galactic halo population and to make comparative analyses between the relative ages of the halo and thick-disk stellar populations. Most recently the isochrones of Bergbusch \\& VandenBerg (2001) have been transformed to the $uvby$ photometric system using the color--$T_{\\rm eff}$ relations of Clem et al.~(2003). Also, the $uvby$--$\\beta$ photometry can provide basic stellar atmospheric parameters as a prelude to detailed chemical abundance studies making use of high-resolution spectroscopy and model atmospheres. Several empirical calibrations already exist in the literature for the conversion of $(b-y)_{\\rm 0}$ or H$\\beta$ to $T_{\\rm eff}$; these calibrations include appropriate metallicity dependences. Index diagrams, such as $c_{\\rm 0}$, $(b-y)_{\\rm 0}$, or the reddening-free $[c_{\\rm 1}]$, $[m_{\\rm 1}]$, or $[c_{\\rm 1}]$, $\\beta$, allow the classification of field stars according to their evolutionary status, permiting us to estimate the stellar surface gravities, also for input into the model-atmosphere analyses. In this paper, $uvby$--$\\beta$ photometry is presented for an additional 411 VMP stars from the HK survey, providing a total database of such photometry for 497 VMP stars, when combined with the data of S96. For the present sample the stars have been selected with [Fe/H] $\\la -1.5$, and 243 were observed in M\\'exico using classical photometric (photoelectric) techniques and 177 in Chile using DFOSC (CCD) techniques. In Sect.~2 the observing and reduction techniques are described briefly, the catalogues of new $uvby$--$\\beta$ data presented, and the $V$ magnitudes and $(b-y)$ colors from the $uvby$ observations compared to magnitudes and $(B-V)$ from the HK survey. In Sect.~3, the photometry is dereddened using a modification of the Schlegel et al.~(1998) reddening maps and also the intrinsic-color calibration of Schuster \\& Nissen (1989); reddenings from the two methods, $E(B-V)$ and $E(b-y)$, are compared. In Sect.~4, [Fe/H] values are derived for the VMP stars using the techniques developed in the HK survey, and probable carbon-enhanced stars are identified based on a comparison of the GP and KP indices. Photometric classifications are derived for the VMP stars in Sect.~5 using the $c_{\\rm 0}$, $(b-y)_{\\rm 0}$ diagram. Stars are found covering a wide range of stellar types from the horizontal branch (HB) to subluminous stars (SL), and from the red giant stars (RG) to the blue horizontal branch (BHB), and other categories include main-sequence (MS), turn-off (TO), subgiant (SG), blue-straggler (BS), and red-horizontal-branch-asymptotic-giant-branch (RHB-AGB) stars. Possible abundance anomalies for some VMP stars have been identified from the $uvby$ photometric indices and diagrams, such as the $[c_{\\rm 1}]$, $[m_{\\rm 1}]$; for example, ten probable Am stars have been found and also a number of possible AGB stars with unusual chemical abundance ratios or binary companions. Distance estimates are made for the VMP stars in Sect.~6 using $uvby$ photometry plus various methods and new calibrations, and also using the $UBV$ photometry and techniques developed in the HK survey. Comparisons of these photometric distances show reasonably good agreement, considering the paucity of calibrating stars and extrapolations required for the more VMP stars. In Sect.~7, the VMP field stars are compared to the globular cluster M92 in the $c_{\\rm 0}$, $(b-y)_{\\rm 0}$ diagram, using the isochrones of Bergbusch \\& VandenBerg (2001), as transformed to $uvby$ by Clem et al.~(2003), to interpolate relative and absolute ages. A number of VMP stars apparently 1--3 Gyrs younger than M92 are noted, and their importance for understanding the formation and evolution of the Galactic halo discussed. ", "conclusions": "\\begin{enumerate} \\item The overall VMP HK-survey sample contains a wide range of stellar types, ranging from horizontal branch stars to subluminous, and from red giant stars to the blue horizontal branch. \\item The dereddened $c_{\\rm 0},(b-y)_{\\rm 0}$ diagram has been shown to be quite useful for providing photometric classifications of the VMP stars analogous to types derived from GC color-magnitude diagrams, such as Turn-Off stars (TO), SubGiants (SG), Red Giants (RG), Horizontal Branch stars (HB), Blue Horizontal Branch stars (BHB), Blue Stragglers (BS), SubLuminous stars (SL), and so forth (see Fig.~6). \\item The intrinsic-color calibration of Schuster \\& Nissen (1989), as modified slightly by Nissen (1994), is shown to provide reddening excesses, $E(b-y)$ or $E(B-V)$, very similar to the adopted reddening estimates derived in this publication from the maps of Schlegel, Finkbeiner, \\& Davis (1998) (see Eq.~1). No significant systematic offsets between these two dereddening techniques are noted (see Fig.~3). \\item A number of VMP stars have been noted with probable anomalous photometric traits, especially from the $m_{\\rm 1}$ and $[m_{\\rm 1}]$ indices; two such groups stand out. First, there are several stars with $(b$--$y)_{\\rm 0} \\la 0\\fm45$ and with $m_{\\rm 0} \\ga 0\\fm17$, much larger than would be expected for VMP stars with [Fe/H] $\\la -1.5$. Most of these have been classified SG, and some show clear evidence of photometric variability. These are perhaps analogous to stars discussed in S96 with larger than expected $[m_{\\rm 1}]$ values. We suggest here that these are misclassified AGB stars with unusual chemical-abundance ratios, photometric variability, and/or binary companions. \\item The second group of anomalous stars are those ten classified BS and having $m_{\\rm 1}$, $[m_{\\rm 1}]$, and $(U-B)_{\\rm 0}$ values indicating nearly solar [Fe/H] values. There is a clear discrepancy here between these photometric indices and the KP index used to derive [Fe/H] for the HK survey. These stars are very similar to the Am stars identified by Wilhelm et al.~(1999a, 1999b) and have been noted as ``BS (Am)'' in Table 5. \\item The photometric distances from the $UBV$ and $uvby$ photometries agree reasonably well considering the problems, lack of calibrating stars, and extrapolations needed for the more VMP stars. Our Hipparcos-based, photometric calibration for $M_{\\rm v}$ seems to work quite well for the turn-off, main-sequence, and subgiant VMP stars, as suggested in Figs.~7 and 8. \\item In the $c_{\\rm 0},(b-y)_{\\rm 0}$ diagram, the youngest VMP stars appear to have ages 1--3 Gyrs younger than the GC M92. Uncertainties in the [Fe/H] scale for M92 would tend to increase this age difference even more. (The interstellar reddening of M92 seems to be well determined but might be as uncertain as $\\pm 0\\fm01$). Such younger VMP stars are showing evidence for important details upon the overall formation and evolution of the Galaxy, such as possible hierarchical star-formation/mass-infall for the VMP material, and/or accretion processes from other (dwarf) galaxies with different formation and chemical-enrichment histories. \\end{enumerate}" }, "0403/astro-ph0403547_arXiv.txt": { "abstract": "{ We considered the structure of steady--state plane--parallel radiative shock waves propagating through the partially ionized hydrogen gas of temperature $T_1 = 3000$~K and density $10^{-12}~\\gcc\\le\\rho_1\\le 10^{-9}~\\gcc$. The upstream Mach numbers range within $6\\le M_1\\le 14$. In frequency intervals of hydrogen lines the radiation field was treated using the transfer equation in the frame of the observer for the moving medium, whereas the continuum radiation was calculated for the static medium. Doppler shifts in Balmer emission lines of the radiation flux emerging from the upstream boundary of the shock wave model were found to be roughly one--third of the shock wave velocity: $-\\dv\\approx \\frac{1}{3}U_1$. The gas emitting the Balmer line radiation is located at the rear of the shock wave in the hydrogen recombination zone where the gas flow velocity in the frame of the observer is approximately one--half of the shock wave velocity: $-V^*\\approx\\frac{1}{2} U_1$. The ratio of the Doppler shift to the gas flow velocity of $\\dv/V^* \\approx 0.7$ results both from the small optical thickness of the shock wave in line frequencies and the anisotropy of the radiation field typical for the slab geometry. In the ambient gas with density of $\\rho_1\\ga 10^{-11}~\\gcc$ the flux in the $\\Ha$ frequency interval reveals the double structure of the profile. A weaker $\\Hb$ profile doubling was found for $\\rho_1\\gtrsim 10^{-10}~\\gcc$ and $U_1\\lesssim 50~\\kms$. The unshifted redward component of the double profile is due to photodeexcitation accompanying the rapid growth of collisional ionization in the narrow layer in front of the discontinuous jump. ", "introduction": "It is now a well--established fact that hydrogen emission lines are a characteristic feature of radially pulsating stars of various types. Strong hydrogen emission is observed in Mira type \\citep{Joy:1947,Joy:1954}, W~Vir \\citep{Abt:1954,Wallerstein:1959} and RV~Tau \\citep{Preston:1962} pulsating variables. Moreover, in the spectra of Mira stars Balmer emission lines persist during the major part of the pulsation period \\citep{Joy:1947,Richter:Wood:2001}. The intensity of the hydrogen emission seems to correlate with the amplitude of the pulsation since RR~Lyr variables exhibit only weak emission lines \\citep{Preston:1964} and in classical Cepheids (for example in $\\beta$~Dor) hydrogen emission lines are scarcely detected \\citep{Hutchinson:1975}. The hydrogen emission in the spectra of pulsating stars is thought to be due to radiative cooling of the gas compressed by the shock wave propagating through the stellar atmosphere in each pulsation cycle \\citep{Kraft:1959,Wallerstein:1959,Abt:Hardie:1960,Gorbatskii:1961}. High resolution spectroscopy reveals the doubling of $\\Ha$ and $\\Hb$ emission profiles, whereas profiles of higher Balmer lines exhibit only the asymmetry. This feature is observed not only in Mira stars \\citep{Bidelman:Ratcliffe:1954,Fox:Wood:Dopita:1984,Gillet:1988,Woodsworth:1995} but also in W~Vir and RV~Tau variables \\citep{Lebre:Gillet:1991,Lebre:Gillet:1992}. \\citet{Bidelman:Ratcliffe:1954} explained the $\\Ha$ profile doubling observed in the Mira type star T~Cen as an absorption reversal rather than a real duplicity, the absorption resulting from the cool hydrogen gas above the propagating shock wave. \\citet{Willson:1976} interpreted the double structure of the emission profiles in terms of a spherically symmetric shock wave with a radial distance from the center of the star appreciably larger than the radius of the photosphere. According to this model the flux of the redshifted component emerges from the back side of the shock wave moving outward from the observer. \\citet{Woodsworth:1995} modelled double $\\Ha$ profiles as three emission components of equal width, two of which are blended. Thus, different phenomenological models demonstrate the ambiguity existing so far in our understanding of the origin of the Balmer emission lines produced by the shock waves in stellar atmospheres. This paper is the fifth in our series on the structure of radiative shock waves. In our previous Papers I--IV \\citep{Fadeyev:Gillet:1998,Fadeyev:Gillet:2000,Fadeyev:Gillet:2001,Fadeyev:2002} we presented the method of computation and described results for the structure of radiative shock waves propagating through partially ionized hydrogen gas with temperature and density typical for atmospheres of pulsating late--type stars. An advantage of this approach is that the gas dynamics, radiation field and atomic level populations are considered self--consistently for the whole shock wave model. However, in all our previous papers the radiative transfer was treated in a static medium approximation, so that we were unable to compare the calculated monochromatic radiation flux with observed emission profiles. Below we describe the shock wave models with Doppler shifts in the line profiles computed from the transfer equation in the frame of the observer. ", "conclusions": "In our attempts to solve the transfer equation for the shock wave structure in the co--moving frame we encountered a severe difficulty because of the numerical instability arising at the velocity discontinuity. In the present study this difficulty could be circumvented because the role of the radiation field in spectral lines is quite small in comparison with that of the continuum. According to our estimates the Doppler shifts in the hydrogen lines do not affect perceptibly either the structure or the radiative losses of the shock wave. In particular, the emergent flux integrated over the line frequency interval was found to be the same within $\\lesssim 1$\\% for both the static and the moving medium. This allowed us to leave out the effects of the Doppler shifts from the global iteration procedure and to solve the transfer equation in the frame of the observer only in the final iteration. The most remarkable result of our study is that the shock wave models show the double emission structure in the $\\Ha$ and $\\Hb$ profiles of the emergent radiation flux which is well known from high resolution high signal--to--noise ratio spectroscopy. We showed that the redward emission feature results from the narrow layer just ahead of the discontinuous jump within which the growth of collisional ionization is accompanied by photodeexcitation onto the second atomic level. The contribution of preshock photodeexcitation decreases with decreasing density of the ambient gas, therefore the double emission feature can be considered as a tool for diagnostics of stellar atmospheres with propagating shock waves. Here one should bear in mind, however, that in the framework of our model the unpertubed gas is at rest with respect to the observer, whereas in pulsating stars the gas ahead of the shock wave falls down onto the star with velocity in the range of one to a few dozen $\\kms$. Thus, the preshock emission feature in $\\Ha$ and $\\Hb$ profiles should be observed as a redshifted component. Another important conclusion is that the velocity inferred from Doppler shifts of Balmer lines is roughly one--third of the shock wave velocity: $\\dv\\approx \\frac{1}{3} U_1$. This is due to the fact that the gas layers emitting the Balmer line radiation are located at the rear of the shock wave in the hydrogen recombination zone where the velocity in the frame of the observer is roughly one half of the shock wave velocity: $-V^*\\approx\\frac{1}{2}U_1$. The ratio of the Doppler shift to the gas flow velocity of $\\dv/V^*\\approx 0.7$ results from the small optical thickness of the shock wave model and the anisotropy of the radiation field produced by the shock wave. $\\Ha$ is the broadest emission line with FWHM comparable to the velocity of the shock wave $U_1$. This, as well as the significant contribution from the preshock photodeexcitation zone make this emission line less appropriate for inferring the velocity of the shock wave from observationally measured Doppler shifts. However the width of the emission profile is proportional to the temperature of the hydrogen atoms behind the discontinuous jump $\\Ta^+$ which is related to the shock wave velocity $U_1$ via the Rankine--Hugoniot relations. Thus, the width of the Balmer line is a function of two general quantities: the shock wave velocity $U_1$ and the ambient gas density $\\rho_1$. In particular, the FWHM of $\\Ha$ decreases by a factor of two with decreasing gas density within $10^{-9}~\\gcc\\le\\rho_1\\le 10^{-12}~\\gcc$. Though our results are consistent with the observations, there is a number of other parameters that determine the propagation of the shock wave in the atmospheres of pulsating stars. In particular, our model is confined to a flat finite slab and therefore does not take into account the cool hydrogen gas of the outer stellar atmosphere above the propagating shock wave. Thus, we cannot exclude the role of absorption in the formation of the double emission structure in Balmer lines. The presence of such absorption in emission profiles observed in RV~Tau and W~Vir stars was pointed out, for example, by \\cite{Lebre:Gillet:1991,Lebre:Gillet:1992}." }, "0403/hep-ph0403134_arXiv.txt": { "abstract": "We show that the non-standard neutrino interactions can play a role as sub-leading effect on the solar neutrino oscillations. We observe that very small flavor universality violations of order of 0.1-0.2 $G_F$ is sufficient to induce two phenomena: suppression of the $\\nu_e$-earth regeneration and a shift of the resonance layer in the sun. We obtain these phenomena even in the absence of any flavor changing interactions. We discuss their consequences and confront with a global analysis of solar+KamLAND results. We conclude that a new compatibility region in the $\\Delta m^2 \\times \\tan^2 \\theta_{\\odot}$ , which we call very low Large Mixing Angle region is found for $\\Delta m^2 \\sim 10^{-5}$ eV$^2$ and $\\tan^2 \\theta_{\\odot}= 0.45$. ", "introduction": "In the last years, the discovery of neutrino oscillation in solar and reactor experiments selected as a more probable explanation to the solar neutrino problem the so called Large Mixing Angle (LMA) MSW solution. The SNO~\\cite{sno,sno-all} and the KamLAND~\\cite{Kam} experiments confirm and refine the trend of the evidences of neutrino oscillations due the solar neutrino observations, as measured by Homestake~\\cite{Cl}, SAGE~\\cite{sage}, GALLEX~\\cite{gallex}, GNO~\\cite{gno} and Super-Kamiokande~\\cite{SK,SK2}. As a result, the solar oscillation parameters have pinned down to $ 6\\times 10^{-5}$ eV$^2 <\\Delta m^2 < 1\\times 10^{-4}$ eV$^2$ and $0.3 < \\tan^2 \\theta_{\\odot} < 0.55$ at $2\\sigma$~\\cite{us:msw2}. Several analyzes have arrived to same conclusions~\\cite{balan,fogli,valle2,alia,crem,choubey}. In a more general context, sub-leading effects can change this picture, which motivate us to investigate the robustness of the determination of the solar parameters. In this letter, we assume that non-standard neutrino interactions, which we parameterized by two parameters $\\epsilon'$ and $\\epsilon$, are present, relaxing the allowed region of the parameters. In the presence of non-standard neutrino interactions, we have found that the allowed interval for $\\Delta m^2$ increases, rescuing the very low LMA region, $\\Delta m^2 \\sim 1\\times 10^{-5}$ eV$^2$, and the high part of LMA region, $\\Delta m^2 \\sim 2\\times 10^{-4}$ eV$^2$, respectively due the suppression of earth matter and due to a $\\Delta m^2$ shift induced by a non-zero $\\epsilon'$. ", "conclusions": "We showed that NSNI will affect the fit in the LMA region of the MSW solution to the solar neutrino anomaly. When one takes into account the KamLAND results, positive values of the $\\epsilon'$ push the allowed region of the neutrino parameters $\\Delta m^2$ and $\\tan^2\\theta_{\\odot}$ at 95\\% C.L. from pure MSW low-LMA and high-LMA to a completely new region in which $\\Delta m^2$ is lower than the previous two ones, which we call very-low-LMA. If one chooses $\\epsilon'<0$, the preferred allowed region tends to higher values of $\\Delta m^2$. Almost all our conclusions below are independent of specific sources of the non-standard neutrino interactions, that could be present in interactions with d-quarks, u-quarks or electrons. We have found that the main effects of the presence of the NSNI interactions are: \\begin{itemize} \\item{displacement of low-LMA region to lower (higher) values of $\\Delta m^2$, for a positive (negative) value of $\\epsilon'$.} \\item{suppression of Earth regeneration at $\\Delta m^2\\sim 10^{-5}$ eV$^2$ for positive values of $\\epsilon'$.} \\item{Due to suppression of Earth regeneration, appearance of a new region of compatibility between solar and KamLAND data around $\\Delta m^2\\sim 10^{-5}$ eV$^2$, with no spectrum distortion for the low-energy SK and SNO data.} \\item{improvement of high-LMA fit quality for positive values of $\\epsilon'$} \\item a 1 kton-yr of KamLAND can make a strong statement about the existence of non-standard neutrino interactions. The striking signal of this NSNI would be the location of the prefered oscillation parameters in the very low or in the high LMA region. \\end{itemize} {\\em Note added:} When we were finishing our paper, an article by Friedland, Lunardini and Pe\\~na-Garay (hep-ph/0402266) appeared, which discusses topics similar to the ones discussed in our paper, where we discuss not only the non-standard neutrino interaction induced by d-quarks case as well the u-quarks and electrons. Also we made a quantitative statement about the role of more statistics on KamLAND experiment, combined with the present solar neutrino data, to put more restrictive bounds on non-standard neutrino interactions." }, "0403/hep-ph0403299_arXiv.txt": { "abstract": "\\noindent We analyze the uncertainties involved in obtaining the injection spectra of UHECR particles in the top-down scenario of their origin. We show that the DGLAP $Q^2$ evolution of fragmentation functions (FF) to $Q=M_X$ (mass of the X particle) from their initial values at low $Q$ is subject to considerable uncertainties. We therefore argue that, for $x\\lsim 0.1$ (the $x$ region of interest for most large $M_X$ values of interest, $x\\equiv 2E/M_X$ being the scaled energy variable), the FF obtained from DGLAP evolution is no more reliable than that provided, for example, by a simple Gaussian form (in the variable $\\ln(1/x)$) obtained under the coherent branching approach to parton shower development process to lowest order in perturbative QCD. Additionally, we find that for $x\\gsim0.1$, the evolution in $Q^2$ of the singlet FF, which determines the injection spectrum, is ``minimal'' --- the singlet FF changes by barely a factor of 2 after evolving it over $\\sim$ 14 orders of magnitude in $Q\\sim M_X$. We, therefore, argue that as long as the measurement of the UHECR spectrum above $\\sim10^{20}\\ev$ is going to remain uncertain by a factor of 2 or larger, it is good enough for most practical purposes to directly use any one of the available initial parametrisations of the FFs in the $x$ region $x\\gsim0.1$ based on low energy data, without evolving them to the requisite $Q^2$ value. ", "introduction": "One of the main problems in understanding the origin of the observed Ultra-High Energy Cosmic Ray (UHECR) events with energy $E\\gsim10^{20}\\ev$\\cite{uhecr_obs} --- below we will sometimes refer to these as Extreme Energy Cosmic Ray (EECR) events --- is the difficulty of producing such enormously energetic particles in astrophysical environments by means of known acceleration mechanisms. There are but a few astrophysical objects --- among which are, perhaps, Gamma Ray Burst (GRB) sources and a class of powerful radio galaxies --- where protons can in principle be accelerated to requisite energies (at source) of $\\gsim10^{21}\\ev$ by the standard diffusive shock acceleration mechanism albeit with optimistic assumptions on the values of the relevant parameters. However, even for these objects, their locations and spatial distributions are not easy to reconcile with the observed spectrum and large-scale isotropy of the UHECR particles. (For recent reviews on astrophysical source origin of EECR see, for example, Refs.~\\cite{springer_book,torres_rev}). \\para An alternative mechanism of producing the EECR particles is provided by the so-called ``top-down'' (TD) scenario (see \\cite{physrep} for a review) in which the EECR particles are envisaged to result from {\\it decay} of some sufficiently massive particles, generically called ``X'' particles, of mass $M_X\\gg10^{20}\\ev$, which could originate from processes in the early Universe. This is in contrast to the conventional ``bottom-up'' scenario in which {\\it all} cosmic ray particles including the EECRs are thought to be produced through processes that accelerate particles from low energies to the requisite high energies in suitable astrophysical environments. \\para The X particles of the TD scenario, if at all they exist in Nature, are most likely to be associated with some kind of new physics at some sufficiently high energy scale that could have been realized in an appropriately early stage of the Universe. Two possibilities for the origin of the X particles have been discussed in the literature: They could be short-lived particles released in the Universe today from cosmic topological defects such as cosmic strings, magnetic monopoles, etc.~\\cite{td_book} formed in a symmetry-breaking phase transition in the early Universe. Alternatively, they could be some metastable (and currently decaying) particle species with lifetime larger than or of the order of the age of the Universe. \\para Since the mass scale $M_X$ of the hypothesized X particle is well above the energy scale currently available in accelerators, its primary decay modes are unknown and likely to involve elementary particles and interactions that belong to unknown physics beyond the Standard Model (SM). However, irrespective of the primary decay products of the X particle, the observed UHECR particles must eventually result largely from ``fragmentation'' of the Standard Model quarks and gluons, that come from the primary decay products of the X particles, into hadrons. The most abundant final observable particle species in the TD scenario are expected to be photons and neutrinos from the decay of the neutral and charged pions, respectively, created in the parton fragmentation process, together with a few percent baryons (nucleons). The injection- or the source spectra of various species of UHECR particles (nucleons, photons and neutrinos) in this TD scenario are thus ultimately determined by the physics of the parton fragmentation process. The final observable UHECR particle spectra are determined by further processing of these injection spectra due to extragalactic and/or Galactic propagation effects depending on where the X particle decay takes place. Clearly, in order to test the predictions of the TD scenario against UHECR experimental data, it is crucial to be able to reliably calculate the injection spectra of various UHECR particles in this scenario. This is the subject we concern ourselves with in this paper. \\para The problem at hand is essentially the same as determining the single-particle inclusive spectrum of hadrons produced, for instance, in the process $e^+e^-\\to \\gamma/Z\\to q\\bar{q}\\to {\\rm hadrons}$ (see, for example, \\cite{ellis_book}). The primary quarks produced in the collision would in general not be on-shell and would have large time-like virtuality $Q\\sim \\sqrt{s}$, the center-of-mass energy of the process. Each quark would, therefore, reduce its virtuality by radiating a gluon, the latter in turn splitting into a $q\\bar{q}$ pair or into two gluons, and so on. This process gives rise to a parton shower whereby at each stage a virtual parton splits into two other partons of reduced virtualities. This process of parton shower development is well-described by perturbative QCD until the virtuality reduces to $Q=Q_{\\rm hadron}\\sim 1\\gev$ when non-perturbative effects come into play binding partons into colorless hadrons. In the end, the link between partons and hadrons is quantitatively described in terms of fragmentation functions (FFs) $D_a^h(x,Q)$, which give the probability that a parton $a$ produced with an initial virtuality $Q=\\sqrt{s}$ produces the hadron $h$ carrying a fraction $x\\equiv 2E/\\sqrt{s}$ of the energy of $a$ ($E$ being the energy of the hadron)\\footnote{At high energies $E$ of our interest throughout this paper we shall assume $E\\simeq p$, the momentum of the particle.}. The final single particle inclusive spectrum of hadrons is given by a convolution of these FFs with the production probabilities of the primary partons (see next section). \\para In the same way, the problem of determining the injection spectrum of UHECR particles from the decay of X particles essentially reduces to determining the FFs $D_a^h(x,M_X)$ for various hadron species $h$ (pions, nucleons) where $a$ represents the primary partons to which the X particle decays. (Actually, in our present case, we will be interested only in the so-called ``singlet'' FF corresponding to a sum over all partons $a$ as explained later). \\para Clearly, the FFs themselves cannot be directly calculated from first principles entirely within perturbative QCD without extra assumptions about the nature of the non-perturbative process of formation of hadrons from partons. Several different approaches have been taken in the recent literature for evaluating the relevant FFs, which are discussed below. \\para In this paper, we critically examine one of the approaches of evaluating the relevant FFs, namely, the DGLAP evolution equation method~\\cite{fodor, sarkar-toldra,barbot-drees,barbot_thesis,aloisio}, that has been widely used in recent calculations of the UHECR injection spectra in the TD scenario. We discuss the inherent uncertainties involved in this approach in calculating the relevant FFs over the ranges of $x$ and $M_X$ of interest. We also compare the FFs so obtained with those given by a simple analytical expression (given by a Gaussian in the variable $\\ln(1/x)$ as discussed later) obtained within the context of an analytical approach, namely, the coherent branching formalism, to lowest order in perturbative QCD~\\cite{ellis_book}, this analytical approach being valid only under ``small\" $x$ and ``large\" $Q$ approximation. We show that except for ``large\" $x\\gsim 0.1$, the uncertainties involved in obtaining the relevant FFs by numerical solution of the DGLAP evolution equation do not allow much significant advantage of using this numerical method over the simple analytical (but approximate) formula for FFs provided by the coherent branching approach. At the same time, we also find that, in the region $x\\gsim 0.1$, the evolution (in $Q$) of the {\\it singlet} FFs (which is what we are interested in) is very little --- the singlet FF changes by only a factor of 2 or so after evolving it over $\\sim$ 14 orders of magnitude in $Q\\sim M_X$. We explain the reason for this, and argue that, as long as the measurement of the EECR spectrum is going to remain uncertain within a factor of 2 or larger (which is likely to be the case in the foreseeable future), it is good enough for most practical purposes to directly use any one of the available parametrisations of the FFs in the $x$ region $x\\gsim0.1$ based on low energy (say at the Z-pole) data from $e^+e^-\\to {\\rm hadrons}$ experiments even without evolving them in $Q$ by means of DGLAP evolution equation. \\para As mentioned above, the X particle decay process may involve particles and interactions belonging to possible new physics beyond SM. Most of the recent studies using DGLAP evolution equation method have been done in the context of a particular model of the possible new physics beyond SM, namely, the Minimal Supersymmetric Standard Model (MSSM). While these studies are certainly useful, there exists, however, no direct evidence yet of Supersymmetry in general and the MSSM in particular. Indeed, the unknown nature of the physics beyond SM introduces additional uncertainties in the whole problem over and above the intrinsic uncertainties associated with the DGLAP evolution method itself which is fundamentally based on standard QCD. In order to analyze these uncertainties associated with the DGLAP evolution method itself, we restrict our analysis here to the standard DGLAP evolution equations for FFs based on QCD. Also, to keep our analysis simple, we shall illustrate our main results by considering the behavior of the FF for only one of the hadron species, namely, pions; our general conclusion, however, apply to nucleons as well as to other mesons like the K meson, too. \\para The rest of this paper is organized as follows: In the following section we set our notations and express the energy spectrum of hadrons resulting from the decay of the X particle in terms of the singlet fragmentation function (FF). In section 3, we review the various methods of evaluating the FF. Our main results are presented and discussed in section 4, and brief conclusions are presented in section 5. ", "conclusions": "In this paper we have analyzed the uncertainties involved in obtaining the injection spectra of UHECR particles in the top-down scenario of their origin. We have demonstrated that evaluating the relevant FFs at the values of $M_X$ and $x$ of interest by evolving them (in $Q=M_X$) from their initial (parametrised) values at low $Q$ by numerically solving the DGLAP evolution equation for FF is subject to considerable uncertainties. Indeed, we find that for $x\\lsim 0.1$ (the $x$ region of interest for most large values of $M_X$ of interest), the FF obtained from DGLAP evolution cannot be said to be any more reliable than that provided by the simple Gaussian form (in the variable $\\xi$) based on coherent branching approach to parton shower development. At the same time, we also find that for $x\\gsim0.1$, the evolution of the singlet FF, which determines the injection spectrum, is ``minimal'' --- the singlet FF changes by barely a factor of 2 after evolving over $\\sim$ 14 orders of magnitude in $Q\\sim M_X$. We, therefore, argue that as long as the measurement of the EECR spectrum is going to remain uncertain by a factor of 2 or larger (which is likely to be the case in the foreseeable future), it is good enough for most practical purposes to directly use any one of the available {\\em initial} parametrisations of the FFs in the $x$ region $x\\gsim0.1$ based on low energy (say at the Z-pole) data from $e^+e^-\\to {\\rm hadrons}$ experiments, without any need for evolving them to the required EECR $Q^2$ value. \\para {\\bf Acknowledgments}\\\\ This work was begun at the Seventh Workshop on High Energy Physics Phenomenology (WHEPP-7) held at Harish-Chandra Research Institute, Allahabad, India, January 4--15, 2002. We thank all the organizers and participants of that Workshop for providing a stimulating workshop environment. The work of PB is partially supported by a NSF US-India cooperative research grant. RB would like to thank D.~Indumathi for useful discussions." }, "0403/astro-ph0403286_arXiv.txt": { "abstract": "s{Recent developments in solar, reactor, and accelerator neutrino physics are reviewed. Implications for neutrino physics, solar physics, nuclear two-body physics, and r-process nucleosynthesis are briefly discussed.} ", "introduction": "Solar neutrino experiments, especially with the announcement of recent results from the Sudbury Neutrino Observatory (SNO) \\cite{Ahmed:2003kj}, have reached the precision stage. An analysis of the data from SNO as well as data from other solar neutrino experiments (Super-Kamiokande [SK] \\cite{Fukuda:2002pe}, Chlorine \\cite{Cleveland:nv}, and Gallium \\cite{Abdurashitov:2002nt,Hampel:1998xg,Altmann:2000ft}), combined with the data from the reactor experiment KAMLAND \\cite{Eguchi:2002dm}, place severe constraints on the neutrino parameters, especially mixing between first and second generations \\cite{Balantekin:2003dc,deHolanda:2003nj,Balantekin:2003jm}. The neutrino parameter space obtained from such a global analysis, including the neutral-current results from the SNO salt phase, is shown in Fig. \\ref{fig:1} \\cite{Balantekin:2003jm}. \\begin{figure} \\includegraphics[scale=0.35]{salt-kl} \\vspace*{0cm} \\caption{ \\label{fig:1} Allowed confidence levels from the joint analysis of all available solar neutrino data (chlorine, average gallium, SNO and SK spectra and SNO salt phase) and KamLAND reactor data The isolines are the ratio of the shifted $^8$B flux to the SSM value. At best fit (marked by a cross) the value of this ratio is determined to be $1.02$ (from Reference 10).} \\end{figure} The mixing angle between first and second generations of the neutrinos dominates the solar neutrino oscillations whereas the mixing angle between second and third generations dominates the oscillations of atmospheric neutrinos. There are several puzzles in the data. Both mixing angles seem to be close to maximum, very unlike the mixing between quarks. Also the third mixing angle, between first and third generations, seems to be very small, even possibly zero. It is especially important to find out if this mixing angle is indeed different from zero since in the mixing matrix it multiplies a CP-violating phase. Such a CP-violation may have far reaching consequences. To explain the baryon excess (over antibaryons) in the Universe, Sakharov pointed out that it may be sufficient to satisfy three conditions: i) Baryon number non-conservation (which is readily satisfied by the grand unified theories), ii) CP-violation, and iii) Non-equilibrium conditions. It is entirely possible that the CP-violation necessary for the baryogenesis is hidden in the neutrino sector. \\begin{figure}[t] \\includegraphics[scale=0.29]{all-2} \\vspace*{0cm} \\caption{ \\label{fig:2} Allowed regions of the neutrino parameter space with solar-density fluctuations when the data from the solar neutrino and KamLAND experiments are used. The SSM density profile of Reference 14 and the correlation length of 10 km are used. The case with no fluctuations ($\\beta=0$) are compared with results obtained with the indicated fractional fluctuation. The shaded area is the 70 \\% confidence level region. 90 \\% (solid line), 95 \\% (dashed line), and 99 \\% (dotted line) confidence levels are also shown (From Reference 15).} \\end{figure} It is worth pointing out that high-precision solar-neutrino data have potential beyond exploring neutrino parameter space. Here we discuss two such applications to solar physics and to nuclear physics. ", "conclusions": "" }, "0403/astro-ph0403553_arXiv.txt": { "abstract": "We report the discovery of a new binary pulsar, PSR~J1829+2456, found during a mid-latitude drift-scan survey with the Arecibo telescope. Our initial timing observations show the 41-ms pulsar to be in a 28-hr, slightly eccentric, binary orbit. The advance of periastron $\\dot{\\omega}=0.28\\pm0.01$~deg yr$^{-1}$ is derived from our timing observations spanning 200 days. Assuming that the advance of periastron is purely relativistic and a reasonable range of neutron star masses for PSR~J1829+2456 we constrain the companion mass to be between 1.22~M$_\\odot$ and 1.38~M$_\\odot$, making it likely to be another neutron star. We also place a firm upper limit on the pulsar mass of 1.38~M$_\\odot$. The expected coalescence time due to gravitational-wave emission is long ($\\sim$~60~Gyr) and this system will not significantly impact upon calculations of merger rates that are relevant to upcoming instruments such as LIGO. ", "introduction": "The first binary pulsar B1913+16 was discovered by Hulse \\& Taylor (1975). This double neutron star (DNS) system with its 7.75-hr orbital period and large eccentricity ($e=0.6$) has since become a wonderful laboratory for testing general relativity in the strong-field regime \\nocite{tw89} (Taylor \\& Weisberg 1989). Perhaps most importantly, it has provided the first evidence for the existence of gravitational radiation \\citep{tw82}. DNS binaries start life as binary systems with main-sequence stars of mass~$>$~6~M$_{\\odot}$. Eventually the more massive of the two stars undergoes a supernova explosion, leaving a neutron star, sometimes seen as a pulsar. As the less-massive star evolves it increases in size until it overfills its Roche lobe. At this point, matter starts to accrete onto the pulsar causing it to spin up to periods as short as a few milliseconds \\nocite{acrs82} (Alpar et al.~1982), a process known as recycling. In most cases the outer layers of the companion are blown away after accretion, exposing the core of the companion and producing a white dwarf-millisecond pulsar binary. For some recycled systems, however, the companion is massive enough to explode as a supernova. In most cases, this violent event will disrupt the binary system. The DNS binaries are those systems fortunate enough to survive. For further details see \\cite{bv91}. Our understanding of the DNS binary population is currently hampered by small-number statistics. Large-scale pulsar surveys are being carried out by a number of groups in order to improve this situation by increasing the sample of objects. In this {\\it Letter} we report on the results of a drift-scan survey using the Arecibo telescope. The survey was conducted at 430 MHz and covered mostly intermediate Galactic latitudes ($|b|<60^{\\circ}$) to optimise the likelihood of finding millisecond and binary pulsars (see \\nocite{cc97} e.g.~Cordes \\& Chernoff 1997). The known pulsars detected are described along with preliminary parameters for a new 41-ms pulsar J1829+2456 which is likely to be a DNS binary. We compare the properties of this new system with the known DNS and other relativistic binaries. Finally, we outline the future observational prospects for this system. ", "conclusions": "\\subsection{Mass determination} The orbital parameters in Table~\\ref{binpars} can be used to constrain the mass of the binary system. The Keplerian mass function relates the orbital period, $P_{b}$, and the projected semi-major axis, $x$, to the masses of the binary components. For this system we calculate a mass function \\begin{equation} f(m_1,m_2) = \\frac{4\\pi^{2} x^3}{P_b^2 T_{\\odot}} = \\frac{(m_{2}\\sin i)^{3}}{(m_{1} + m_{2})^{2}} = 0.294\\ \\rm{M}_{\\odot}, \\label{cmass} \\end{equation} where $T_{\\odot} = GM_{\\odot}c^{-3} = 4.925490947 \\mu$s, $i$ is the inclination between the plane of the orbit and the line of sight, $x$ is in light-seconds, $P_b$ is in seconds and the pulsar and companion masses $m_1$ and $m_2$ are in Solar masses. Neutron stars are observed to have a narrow range of masses. Given the data in \\cite{tc99}, and the low measured mass of PSR~J0737$-$3039B (Lyne et al. 2004), we expect the pulsar mass to lie between 1.25~M$_\\odot$ and 1.47~M$_\\odot$. For an orbit viewed edge-on ($i=90^{\\circ}$), and a minimum neutron star mass for PSR~J1829+2456 of $m_{1}=1.25\\ \\rm{M}_\\odot$, we calculate a minimum companion mass $m_2 \\simeq$ 1.22 M$_\\odot$. As shown in the following equation, the measurement of the advance of periastron, $\\dot{\\omega}$, if assumed to be purely relativistic, allows a measurement of the sum of the masses: \\begin{equation} \\dot\\omega=3\\left(\\frac{2\\pi}{P_{b}}\\right)^{5/3}T_{\\odot}^{2/3}(m_{1}+m_{2})^{2/3}(1-e^{2})^{-1}. \\label{omegadot} \\end{equation} When the measured $\\dot \\omega$ is used in combination with the constraint that $\\sin i<1$, we obtain an upper limit on the maximum pulsar mass~$m_1<1.38$ M$_{\\odot}$. Similarly, assuming $m_1>1.25$ M$_{\\odot}$, the maximum value of $\\dot{\\omega}$ implies $i>66^{\\circ}$. These constraints are shown in a mass-mass diagram in Fig.~\\ref{massdiag}. Although we cannot currently rule out a massive white dwarf or main-sequence star, given the similar mass functions, spin and orbital parameters of J1829+2456 to other DNS binaries (see Table~\\ref{DNSsystems}) it seems most likely that the companion is another neutron star. \\begin{figure} \\includegraphics[width=8cm, angle=0]{fig3.ps} \\caption{The constraints on the masses of the pulsar and companion. The vertical lines are the mass limits for neutron stars described in Thorsett \\& Chakrabarty (1999), and the low measured mass of PSR~J0737$-$3039B (Lyne et al. 2004). The lower companion mass limit is given by the mass function assuming a maximum inclination of $90^{\\circ}$. The advance of periastron, $\\dot{\\omega}$, if assumed to be purely relativistic, provides the constraint for the maximum companion mass and maximum pulsar mass.} \\label{massdiag} \\end{figure} \\subsection{A search for the companion} The accuracy of the position of PSR~J1829+2456 inferred through radio timing is sufficient to allow for a search for an optical companion. No optical companion is present in the uncalibrated plates of the Digitised Sky Survey. Preparations are underway for a deep optical search for the companion star. Given the possibility that the companion of J1829+2456 is a neutron star, a search was made of the available radio data for a periodic signal which could be coming from the companion if it were active as a radio pulsar. This search was given greater significance following the recent detection of 2.8-s pulsations from the companion of PSR~J0737$-$3039A (Lyne et \\nocite{lbk+04} al.~2004). Initial searches of the data (at the DM of PSR~J1829+2456) looking for other periodic signals showed that, if the companion was an active radio pulsar, it was too weak to be easily detectable. To carry out a more sensitive search for the companion, we made use of the fact that the orbital parameters of the system are known and that the pulsations from the companion star would be Doppler shifted in the opposite sense to PSR~J1829+2456 due to its orbital motion about their common centre of mass. To correct for this effect, we applied the first-order Doppler formula to calculate the effective sampling interval in the rest-frame of the companion: \\begin{equation} t_{\\rm companion} = t_{\\rm samp} (1 + v(t)/c), \\end{equation} where $t_{\\rm samp}$ is the sampling interval at the observatory and $v(t)$ is the effective line-of-sight radial velocity of the companion as a function of time, $t$. We used TEMPO to calculate $v(t)$ to account for the effects of orbital motion in the binary system (using the orbital parameters given in Table~\\ref{binpars}, assuming a pulsar mass of 1.35~M$_\\odot$ and changing only the longitude of periastron by $180^{\\circ}$) as well as contributions from the Earth's rotation and motion about the Sun. All available data were dedispersed and resampled in this way before being passed through the standard search analysis. No significant candidates were found down to a S/N of 5. This corresponds to a 430-MHz flux limit of $\\sim$~0.1~mJy. While the lack of detection of the companion suggests that it is too weak to be seen as a radio pulsar, or unfavourably beamed, it should be noted that the companion to PSR~J0737$-$3039A is only clearly detectable over certain phases of the orbit (Lyne et al.~2004). If any companion to PSR~J1829+2456 was behaving in a similar fashion it may not be visible in the current data set. Further, more sensitive, searches will be carried out as more data are taken. \\subsection{Expected relativistic parameters} Using the constrained range of masses for the pulsar and companion the expected orbital period derivative and period for geodetic precession due to misalignment of the spin axis of the pulsar with the orbital angular momentum vector \\citep{bo75b} can be determined. Assuming a combined mass of 2.53~M$_\\odot$, the expected values for $\\dot P_{b}$ and the expected timescale for geodetic precession are given in Table~\\ref{binpars}. The expected $\\gamma$ should be measurable after 18 months of timing data. The expected $\\dot P_{b}$ is too small to be measurable without a long-term (5--10 yr) timing campaign. The expected timescale for geodetic precession is $\\sim$~4200~yrs, too long to be measurable through pulse profile or timing variations. \\begin{table} \\caption{Measured, derived, and expected parameters for PSR~J1829+2456 from 29 TOAs from MJDs 52786 to 52986.} \\begin{center} \\begin{tabular}{l l} \\hline \\multicolumn{2}{c}{Measured Parameters} \\\\ \\hline Right ascension (J2000) (h:m:s) & 18:29:34.6 (1) \\\\ Declination (J2000) ($^{\\circ}$:':'') & 24:56:19 (2) \\\\ Period (ms) & 41.00982358 (1)\\\\ Epoch of period (MJD) & 52887 \\\\ Projected semi-major axis (s) & 7.2360 (1) \\\\ Dispersion measure (pc cm$^{-3}$) & 13.9 (5) \\\\ Binary period (days) & 1.176028 (1) \\\\ Eccentricity & 0.13914 (4) \\\\ Longitude of periastron (deg) & 229.94 (1) \\\\ Epoch of periastron (MJD) & 52848.57977 (3)\\\\ 430-MHz flux density (mJy) & 0.3 (1) \\\\ Pulse width at 50\\% of peak $w_{50}$ (ms) & 1.07 (16) \\\\ Pulse width at 10\\% of peak $w_{10}$ (ms) & 2.07 (16) \\\\ Rate of advance of periastron (deg yr$^{-1}$) & 0.28 (1) \\\\ Mass function (M$_\\odot$) & 0.29413 (1) \\\\ RMS residual to fit ($\\mu$s) & 19 \\\\ \\hline \\multicolumn{2}{c}{Derived Parameters} \\\\ \\hline Galactic longitude (J2000) (deg) & 53.343 (1) \\\\ Galactic latitude (J2000) (deg) & 15.612 (1) \\\\ Distance (kpc)$^{a}$ & $\\sim$ 1.2 \\\\ Mean orbital speed (km s$^{-1}$) & 134 \\\\ Total system mass (M$_\\odot$) & 2.5 (2) \\\\ Minimum companion mass (M$_\\odot$) & 1.22 \\\\ Maximum pulsar mass (M$_\\odot$) & 1.38 \\\\ \\hline \\multicolumn{2}{c}{Expected Relativistic Parameters} \\\\ \\hline Orbital period derivative ($\\times10^{-12}$) & $-$0.02 \\\\ Coalescence time (Gyr) & 60 \\\\ Geodetic precession rate (deg yr$^{-1}$) & 0.075 \\\\ Relativistic time-dilation\\\\ and gravitational redshift $\\gamma$ (ms) & 1.3 \\\\ \\hline \\end{tabular} \\end{center} a: Distance inferred from DM \\citep{cl02}.\\\\ The numbers in parentheses are the 1--$\\sigma$ uncertainties in the least significant digit quoted. \\label{binpars} \\end{table} \\subsection{Comparison with other relativistic binary systems} If our future measurements show PSR~J1829+2456 to be a double neutron star system it will be the seventh such system to be discovered. Table~\\ref{DNSsystems} contains a list of the most relativistic binary pulsar systems and their orbital parameters. We also list the coalescence time $\\tau_{\\rm GW}$ due to the emission of gravitational radiation of each system. Following Lorimer (2001)\\footnote{We reproduce the formula here to correct a typographic error in the original citation.}, we can approximate the detailed calculations of $\\tau_{\\rm GW}$ by Peters \\nocite{pet64} (1964) via \\begin{equation} \\label{coaltime} \\tau_{\\rm GW} \\simeq 10^7 \\, {\\rm yr} \\, \\left(\\frac{P_{b}}{\\rm hr}\\right)^{8/3} \\left(\\frac{\\mu}{{\\rm M}_{\\odot}}\\right)^{-1} \\left(\\frac{m_1+m_2}{{\\rm M}_{\\odot}}\\right)^{-2/3} (1-e^{2})^{7/2}, \\end{equation} where the reduced mass $\\mu = m_1 m_2 / (m_1+m_2)$. For any reasonable range of $m_1$ and $m_2$ discussed above, we find $\\tau_{\\rm GW} \\sim$~60~Gyr. The systems listed in Table~\\ref{DNSsystems} span a wide range of orbital parameters but can be broadly split into systems that will and will not coalesce within a Hubble time, with PSR~J1829+2456 lying toward the edge of the non-coalescing group. The relatively long coalescence time for J1829+2456 means that it will not affect the DNS merger rate calculations (e.g. Kalogera et al. 2004) for gravitational wave detectors such as LIGO. \\begin{table*} \\caption{The orbital parameters of various eccentric binary systems.} \\begin{center} \\begin{tabular}{l r@{.}l r@{.}l r@{.}l r@{.}l r@{.}l r@{.}l c r@{.}l c c} \\hline PSR &\\multicolumn{2}{c}{$P$} &\\multicolumn{2}{c}{$P_{b}$} &\\multicolumn{2}{c}{$a_{1}\\sin~i$} &\\multicolumn{2}{c}{$e$} &\\multicolumn{2}{c}{$\\dot \\omega$} &\\multicolumn{2}{c}{$\\dot P _{b}$} & $f(m)$ &\\multicolumn{2}{c}{$m_{1} + m_{2}$} &$\\tau_{\\rm GW}^{\\dag}$ & References\\\\ &\\multicolumn{2}{c}{(ms)} &\\multicolumn{2}{c}{(days)} &\\multicolumn{2}{c}{(lt-s)} &\\multicolumn{2}{c}{} &\\multicolumn{2}{c}{(deg yr$^{-1}$)} &\\multicolumn{2}{c}{($\\times 10^{-12}$)} & \\multicolumn{2}{c}{(M$_\\odot$)} &(M$_\\odot$) &(Gyr) & \\\\ \\hline &\\multicolumn{17}{c}{Double neutron star binaries}\\\\ \\hline B1913+16 & 59&03 & 0&323 & 2&34 & 0&617 & 4&227 & $-$2&428 & 0.13 & 2&83 & 0.31 & 1 \\\\ B1534+12 & 37&90 & 0&421 & 3&73 & 0&274 & 1&756 & $-$0&138 & 0.31 & 2&75 & 2.69 & 2 \\\\ B2127+11C & 30&53 & 0&335 & 2&52 & 0&681 & 4&457 & $-$3&937 & 0.15 & 2&71 & 0.22 & 3 \\\\ J1518+4904 & 40&93 & 8&634 & 20&04 & 0&249 & 0&011 &\\multicolumn{2}{c}{--} & 0.12 & 2&62 & 9600 & 4 \\\\ J1811$-$1736 & 104&18 & 18&779 & 34&78 & 0&828 & 0&009 &\\multicolumn{2}{c}{$<$30} & 0.13 & 2&6 & 1700 & 5 \\\\ J0737$-$3039A& 22&70 & 0&102 & 1&42 & 0&088 & 16&88 & $-$1&24$^{\\ast}$ & 0.29 & 2&58 & 0.087 & 6 \\\\ \\hline &\\multicolumn{17}{c}{White dwarf binaries}\\\\ \\hline B2303+46 &1066&37 & 12&34 & 32&69 & 0&66 & 0&010 &\\multicolumn{2}{c}{--} & 0.25 & 2&53 & 4500 & 7 \\\\ J1141$-$6545 & 393&90 & 0&20 & 1&86 & 0&17 & 5&33 &\\multicolumn{2}{c}{$<$50} & 0.18 & 2&30 & 0.59 & 8 \\\\ \\hline &\\multicolumn{17}{c}{Unknown companion}\\\\ \\hline B1820$-$11 & 279&83 &357&76 &200&67 & 0&79 &\\multicolumn{2}{c}{$<10^{-4}$} &\\multicolumn{2}{c}{--} & 0.07 & \\multicolumn{2}{l}{--} & -- & 9 \\\\ J1829+2456 & 41&00 & 1&17 & 7&24 & 0&14 & 0&28 & $-$0&02$^{\\ast}$ & 0.29 & 2&53 & 60 & -- \\\\ \\hline \\end{tabular} \\begin{tabular}{l l} 1: \\cite{ht75a, wt03a, tw89} &6: \\cite{bdp+03} \\\\ 2: \\cite{wol90z, sttw02} &7: \\cite{tamt93, arz95} \\\\ 3: \\cite{agk+90, and92} &8: \\cite{klm+00a} \\\\ \\cite{pakw91, dk96} &9: \\cite{lm89, tamt93} \\\\ 4: \\cite{nst96, hlk+03} &$\\ast$: Predicted value. \\\\ 5: \\cite{lcm+00} &$\\dag$: Calculated using formula \\ref{coaltime}.\\\\ \\end{tabular} \\end{center} \\label{DNSsystems} \\end{table*}" }, "0403/astro-ph0403079_arXiv.txt": { "abstract": "I review cosmological simulations of X-ray clusters. Simulations have increased in resolution dramatically and the effects of radiative cooling, star formation feedback, and chemical enrichment on the ICM are being simulated. The structure and evolution of non-radiatve X-ray clusters is now well characterized. Such models fail to reproduce the observed $L_x-T$ relation, implying the need for additional physics. Simulations adding radiative cooling produce too much cool gas and unreasonably high X-ray luminosities. Simulations including star formation and feedback appear more promising, but need further refinement. New observations should help in this regard. ", "introduction": "As the largest gravitationally bound objects in the universe, clusters of galaxies have attracted the attention of observers and numerical simulators alike. For over a decade, beginning the with pioneering hydrodynamic simulations of Evrard (1990), numerical simulations have been used to understand the physics of X-ray cluster formation and to predict their abundance at high redshift which is a sensitive probe of cosmology (see review Henry in these proceedings.) Observationally, clusters of galaxies have historically been studied in the optical and X-ray portions of the EM spectrum (Forman \\& Jones 1982). X-rays in particular provide an unambiguous method for detecting clusters at low and intermediate redshift, and many surveys have been conducted (Henry, these proceedings.) A number of groups have simulated the formation of statistical ensembles of X-ray clusters (Kang et al. 1994; Bryan et al. 1994a,b; Bryan \\& Norman 1998; Eke, Navarro \\& Frenk 1998; Yoshikawa, Jing \\& Sato 2000) in order to provide a theoretical bridge between what is observed--the X-ray luminosity function (XLF) and X-ray temperature function (XTF)--and the cluster mass function (CMF). The CMF in turn is directly related to the matter fluctuation power spectrum $P(k)$--one of the holy grails of observational cosmology. In so doing, simulators have discovered that X-ray clusters are not the simple gas-bags they were once thought to be. It has been found that quite high resolution is required to converge on the predicted properties of non-radiative clusters (Anninos \\& Norman 1996; Frenk et al. 1999), and that the inclusion of radiative cooling and other non-adiabatic effects strongly affects the clusters' emission and structural properties (e.g., Pearce et al. 2000). This review will follow the development of cosmological simulations of X-ray clusters primarily from a historical perspective, starting with the non-radiative simulations of Evrard (1990) and others and concluding with current models incorporating cooling, star formation, supernova feedback and chemical enrichment. The field has been enlivened by the arrival of new observations and new questions. I will attempt to keep the questions at the forefront of this review, for while some have been convincingly answered, many are still open. Simulations of cluster mergers done outside the framework of CDM-driven structure formation are not reviewed here for space reasons. ", "conclusions": "" }, "0403/astro-ph0403623_arXiv.txt": { "abstract": "s{The Combined EIS-NVSS Survey Of Radio Sources (CENSORS) has been produced with the primary goal of investigating the cosmological evolution of the radio luminosity function. This 1.4GHz sample, complete to the $7.2$mJy level, contains 150 radio sources. Host galaxies are almost entirely identified in optical and near-IR bands and the sample is now approaching 70\\% spectroscopic completeness. We show preliminary results demonstrating how CENSORS will improve upon previous work and how it is applicable to other projects.} ", "introduction": "Radio galaxies and radio loud quasars are among the most powerful objects in the Universe. Observable at high redshifts, they trace large scale structure and are associated with the most massive black holes[\\refcite{rad_bhmass}], they therefore have great potential for revealing the evolution of massive galaxies and investigating the relation between active galactic nuclei (AGN) and galaxy formation. However there are still many questions facing radio-loud AGN astrophysicists. For example, the high redshift cosmological evolution of strongly radio active sources is a property that has not been resolved and so provides a route by which we may gain a better understanding of these objects. This is of far-reaching importance: given their association with the most massive black holes, we would also learn about the evolution of the upper part of the black hole mass distribution. In 1990 Dunlop and Peacock[\\refcite{dp90}] undertook a study of cosmic evolution of the radio luminosity function (RLF) presenting the first evidence of a decline in the comoving number density of powerful radio sources beyond z $\\squig$ ~2.5 (the redshift cut-off). Since then numerous advances have provided a good consensus in the determination of the low redshift RLF (eg.[\\refcite{mob1999}]), but the high redshift evolution and the reality of the redshift cut-off in the radio source population remain areas of controversy. For example, the deep sample of Waddington et al.[\\refcite{wad2001}] shows evidence of a cut-off in number density of low luminosity sources beyond z $>$ 2, but has insufficient sky coverage to investigate the most luminous sources. Conversely, the sample of Jarvis et al.[\\refcite{jarvis2001}] proved too shallow to resolve the issue. Shaver et al.[\\refcite{shaver}] claimed evidence for a sharp decline in number density of flat spectrum sources between $z \\sim 2.5$ and $z \\sim 5$. However Jarvis and Rawlings[\\refcite{jar_raw}] showed that this result was erroneous due to a failure to properly account for the spectral index of the sources. So high redshift RLF evolution of radio sources remains ill-understood. ", "conclusions": "" }, "0403/astro-ph0403309_arXiv.txt": { "abstract": "{ Four Bok globules were studied in the Near-Infrared, through narrow-band filters, centered at the 1.644~$\\mu$m line of [FeII], the H$_2$-line at 2.122~$\\mu$m, and the adjacent continuum. \\hfill\\break\\noindent We report the discovery of \\FE\\ and H$_2$ protostellar jets and knots in the globules CB3 and CB230. The [FeII]-jet in CB230 is defined by a continuous elongated emission feature, superimposed on which two knots are seen; the brighter one lies at the tip of the jet. The jet is oriented in the same direction as the large-scale CO outflow, and emerges from the nebulosity in which a Young Stellar Object is embedded. The H$_2$ emission associated with this jet is fainter and wider than the [FeII] emission, and is likely coming from the walls of the jet-channel. \\hfill\\break\\noindent In CB3 four H$_2$ emission knots are found, all towards the blue-shifted lobe of the large-scale outflow. There is a good correspondence between the location of the knots and the blue-shifted SiO(5$-$4) emission, confirming that SiO emission is tracing the jet-like flow rather well. \\noindent No line emission is found in the other two targets, CB188 and CB205, although in CB205 faint line emission may have been hidden in the diffuse nebulosity near the IRAS position. Around this position a small group of ($\\geq 10$) stars is found, embedded in the nebula. A diffuse jet-like feature near this group, previously reported in the literature, has been resolved into individual stars. ", "introduction": "\\label{int} Outflow and infall are inextricably associated with the very earliest stages of star formation. Even while still accreting matter, a newborn star generates a fast, well-collimated stellar wind that forms jets which sweep up the ambient molecular gas, creating bipolar molecular outflows. The high-velocity winds create shocks, which heat the gas to thousands of K when breaking up into the ambient gas. Understanding the details of the mechanism producing the acceleration of the outflow is fundamental to the understanding of the star formation process itself. \\noindent Emission of millimetre molecular lines (usually CO) at excitation temperatures T$_{\\rm ex} \\approx 10-20$~K, is used to study the outflow's large-scale morphology. Whereas this component consists mostly of swept-up cloud material, and as such offers a time-averaged picture, the {\\it present} flow activity is represented by a fast, hot component, traced by H$_2$ ro-vibrational lines at $T_{\\rm ex} \\sim 2000$~K. Where this jet-like component interacts with the slower flowing gas or the ambient cloud, bow-shaped shock fronts are visible. This component, at T$_{\\rm ex} \\sim 100$~K, can be traced by mm-emission of molecular species that are produced only in a shock-driven chemistry (e.g Bachiller~\\cite{bach}). The link between the 10-100~K gas and the hot jet-component is not well understood. In particular, it is still an open question how the efficiency of the processes leading to chemical anomalies depends on the shock-type (J/C) and -characteristics. The comparison between the shocked gas components at different temperatures is thus fundamental to the study of the jet/outflow system and the energetics of the star forming process. \\begin{figure} \\resizebox{7.5cm}{!}{\\rotatebox{270}{\\includegraphics{0931fig1.eps}}} \\caption[]{TNG/NICS observations of CB3: The H$_2$ emission (still including the contribution from the continuum). The 4 knots (indicated by arrows) of H$_2$ line-emission are not at all visible in the K-continuum image (not shown). } \\label{cb3h2kcont} \\end{figure} \\begin{figure} \\resizebox{8.5cm}{!}{\\rotatebox{270}{ \\includegraphics{0931fig2.eps}}} \\caption[]{{\\bf a}\\ A zoom-in on the four H$_2$ knots in CB3. To better outline k2 and k3, contours for them are drawn with smaller increments than for the other knots in the panel; lowest contour is at $\\sim 2.5\\sigma$. The dashed lines indicate the alignment of pairs of knots with the outflow centre. The centre of the outflow (Codella \\& Bachiller~\\cite{codbach}) lies at offset (0,0) and is indicated by a cross. The 1.3-mm peak (Launhardt \\& Henning~\\cite{launhen}) is indicated by the star. {\\bf b}\\ As a, but overlaid with the contours of integrated SiO(5--4) emission at $-42.5 (\\pm 0.5)$~km s$^{-1}$ (Codella \\& Bachiller~\\cite{codbach}). The angular resolution of the SiO map is 11$\\arcsec$. } \\label{cb3h2} \\end{figure} \\smallskip\\noindent A very good place to study the jet-component, the large-scale outflow, and their interaction are the Bok globules: cold (10 K) and relatively isolated molecular clouds associated with star formation. A catalogue of such objects (at $\\delta > -30^{\\circ}$) was compiled by Clemens \\& Barvainis (\\cite{clemens}; hereafter CB). Because of relatively simple structure, globules form mainly low-mass stars in small numbers, and are therefore without the observational confusion that one encounters in regions like Orion or Ophiuchus. With the italian TNG (Telescopio Nazionale Galileo) we have searched in four globules for the jet-component in the Near-InfraRed (NIR) through narrow-band H$_2$ (2.122~$\\mu$m) and [FeII] (1.644~$\\mu$m) filters. These two lines are particularly useful, as [FeII] traces (J-) shocks with velocities of a few 100~km s$^{-1}$ and is therefore expected to outline the inner jet-channel, closest to the driving source of the flow. Molecular hydrogen, which dissociates at shock velocities $>25-45$~km\\,s$^{-1}$, traces slower (C-) shocks and is excited in bow shocks and in shock-wakes (e.g. Allen \\& Burton~\\cite{allen}) and is therefore a good probe for the region of interaction between jet and ambient material. All four globules have been observed in the broad-band NIR (Yun \\& Clemens~\\cite{yc94a}) and detected at 1.3~mm continuum (Launhardt \\& Henning~\\cite{launhen}) and were found to contain embedded objects. Molecular outflows were also found in all four objects (Yun \\& Clemens~\\cite{yc94b}; Codella \\& Bachiller~\\cite{codbach}). Together, these findings identify these globules as good candidates in a search for protostellar jets. In this paper we report the detection of H2 and [FeII] line emission knots and jets in two of the four globules. ", "conclusions": "\\label{res} \\subsection{Removing the continuum} Observations in the H$_2$ and [FeII] filters contain, apart from the emission in the lines themselves (if present) also a contribution from the K- (in the case of H$_2$) and H-continuum (in the case of [FeII]). Therefore in order to detect the line emission the continuum has to be subtracted. Because the central wavelength of the continuum filters is different from those of the line filters, a proper subtraction requires a careful scaling of the emission detected in the continuum filters. Field stars are used for this, because their emission is expected to be continuum only. \\noindent We have assumed that the wavelength-dependence of the emission in the NIR is proportional to $\\lambda^x$. The value of the slope $x$ is derived from a comparison of the flux of a number of field stars in the line- and continuum filters. \\noindent In CB188 and CB205, no jet-like features are detected after subtraction, whereas in CB3 and CB230 pure H$_2$ and [FeII] emission knots and protostellar jets have been clearly revealed. \\hfill\\break\\noindent Photometry of the line-emission knots has been performed on the subtracted images using the task {\\tt POLYPHOT} in IRAF, roughly including the emission down to a $\\sim 1 \\sigma$ limit. The sensitivity limit (of integrated line flux) is $2-3 \\times 10^{-16}$ erg cm$^{-2}$ s$^{-1}$ arcsec$^{-1}$, both in the H$_{2}$ and in the [FeII] images; the limiting flux increases towards regions of diffuse emission. \\begin{figure*} \\resizebox{10cm}{!}{\\rotatebox{270}{\\includegraphics{0931fg4a.eps}}} \\resizebox{6cm}{!}{\\rotatebox{270}{\\includegraphics{0931fg4b.eps}}} \\caption[]{ Results of the TNG/NICS observations of CB230 of the H$_2$ and [FeII]-emission (still including continuum emission): (Central panel) A zoom-in on the objects detected in H$_2$ in the core of CB230, with contours to better outline the more intense emission. The YSO is indicated by ``A''. (Left) A close-up of the smaller emission object, which is clearly seen to contain two sources (B1 and B2). The right-hand panel shows the contoured [FeII] emission (still including the continuum). A clear difference is seen in the diffuse emission just North of the YSO, with respect to the H$_2$ image (central panel); this is the signature of a jet (see Fig.~\\ref{feiih2closeup}). } \\label{cb230h2} \\end{figure*} \\subsection{CB3} This globule is associated with different generations of star formation, as pointed out by the presence of a NIR Young Stellar Object (YSO; Yun \\& Clemens~\\cite{yc95}, \\cite{yc94a}), a 1.3 mm object (Launhardt \\& Henning \\cite{launhen}), and a sub-mm source (Huard et al. \\cite{huard}). This globule also contains an IRAS point source (IRAS00259+5625), which lies among these objects; its flux is probably derived (in part) from contributions by these various objects. A molecular outflow was detected by Yun \\& Clemens (\\cite{yc92}, \\cite{yc94b}) in CO and mapped in various molecular lines by Codella \\& Bachiller (\\cite{codbach}), who found that the outflow is centered at or very near the 1.3~mm source, which is 17$\\farcs$4 W, 1$\\arcsec$ S of the IRAS position. \\noindent Our NIR images are centered on the centre position of the outflow. Fig.~\\ref{cb3h2kcont} shows the image obtained in the H$_2$ filter. Even though this image still contains a contribution from the continuum, four regions of H$_2$ emission, hereafter called knots k1, k2, k3, and k4 (North to South), are clearly distinguishable; they are not present in the K$_{\\rm cont}$-image (not shown). Fig.~\\ref{cb3h2}a shows a close-up of the central region of the narrow-band line image; with the dashed lines we have tried to suggest how the spots can be aligned in pairs and how they may be traced back to the centre of the outflow (which, within the accuracy of the positions is presumably coincident with the 1.3~mm source). The lines show that the centers of k1 and k3 are aligned with the center of the outflow, while the line connecting the origin of the flow and the centers of k2 and k4 is slightly offset from that. This may suggest a change of direction of the outflow axis with time. \\noindent Though unbeknownst to us at the time of the observations, the features k1 and k4 in Fig.~\\ref{cb3h2}a were already noted in deep NIR images of this region by Launhardt et al. (\\cite{launISO}), who remarked that these two diffuse, non-stellar objects ``might have H$_2$-line emission contributing to their K-band luminosity''. The fact that these features are not visible in our K$_{\\rm cont}$ image confirms that they arise from pure H$_2$ line emission. The absolute coordinates of the four H$_2$ knots, as well as their integrated flux, are given in Table~\\ref{knotparams}. \\noindent Knot k4 lies at a distance of $\\sim 40\\arcsec$ from the centre of the outflow; for an assumed distance of 2.5~kpc (Launhardt \\& Henning~\\cite{launhen}) this corresponds to $\\sim 0.5$~pc. Knot k1 is the nearest to the outflow centre, at a distance of 10$\\arcsec \\approx 0.12$~pc. All knots are found South of the outflow centre, towards the blue lobe of the molecular outflow (Codella \\& Bachiller \\cite{codbach}). The fast outflow component is usually traced by H$_2$ line emission, which originates behind the shocks created where this jet-component interacts with the ambient cloud material. That we do not see any H$_2$ spots associated with the red (northern) lobe of the outflow is likely because it is much more embedded in the globule, and the emission is too much obscured to be visible even at the NIR wavelength of H$_2$. \\noindent Closer inspection of the brightest knots k1 and k4 shows that the emission contours are slightly convex. Together with the knot alignment, and the location of the knots along the outflow axis, this argues in favour of the bow shock model for the interaction of a wind from the YSO and the ambient medium. \\noindent Downstream from the shocks caused by the interaction of the YSO-wind and the ambient medium, in high-density and -temperature gas, one expects to find an enhancement of several molecular species, which are liberated from the mantles of dust grains and injected into the gas phase through endothermic or gas-grain reactions as well as through sputtering. Codella \\& Bachiller (\\cite{codbach}) have shown that the high-velocity component of the outflow in CB3 is well-traced by SiO, which is enhanced in the presence of shocks and which is only present along the main flow axis. In Fig.~\\ref{cb3h2}b we compare the location of the H$_2$ spots with (blue-shifted) channel maps of the emission of SiO(5--4). There is a close correspondence between the location of (at least two of) the H$_2$ spots and the peaks of the SiO distribution. This confirms that in CB3 the SiO emission is tracing the jet-like flow rather well, and indicates that along the outflow axis different temperature regimes coexist: the warm component at $\\sim$100~K (Codella \\& Bachiller 1999) traced by SiO and a hot ($\\ge$2000~K) component revealed by H$_2$. \\noindent In Fig.~\\ref{cb3allobjects} we show the locations of the various tracers of star formation found in this globule, projected on the narrow-band H$_2$ image. An evolutionary sequence can be distinguished, in that we encounter progressively younger objects when going from the NIR-detected object YC1 (Yun \\& Clemens~\\cite{yc95}), via the sub-mm peaks (Huard et al. \\cite{huard}), to the 1.3~mm peak (Launhardt \\& Henning~\\cite{launhen}), which is probably coincident with the centre of the outflow (Codella \\& Bachiller~\\cite{codbach}). This implies that star formation in this globule has been going on for some time. Assuming the IRAS source and the sub-mm peaks refer to the same object, the mutual distance between the 3 star formation indicators in Fig.~\\ref{cb3allobjects} is about 0.2~pc. \\noindent We take this opportunity to point out some inconsistencies in the original papers regarding the relative positions of the various objects plotted in Fig.~\\ref{cb3allobjects}. First, we note that the object identified as YSO (based on its NIR excess) in Yun \\& Clemens (\\cite{yc95}) is different from the one they identified in CB3 in Yun \\& Clemens~\\cite{yc94a}, which is about 100$\\arcsec$ to the North. And even so, although YC1-I is a bright source (K = 8.24~mag.; Yun \\& Clemens~\\cite{yc95}), it does not coincide with any of the sources in our image -- the nearest object (see Fig.~\\ref{cb3allobjects}) is $\\sim 4\\arcsec$ to the South, much more than the estimated positional accuracy of both our objects and YC1-I ($\\sim 2\\arcsec$). \\hfill\\break\\noindent The relative positions of the Huard et al. (\\cite{huard}) sub-mm peak and the IRAS source are incorrect in their Fig.~1: based on the coordinates given by Huard et al. the sub-mm peak should be much closer to the IRAS position ($\\sim 5\\arcsec$ rather than 15$\\arcsec$). Although Launhardt \\& Henning (\\cite{launhen}) do not give a positional accuracy for their 1.3~mm peak, it is probably similar to that of the sub-mm peak (i.e. $\\sim 3\\arcsec$), suggesting that these are different objects. \\subsection{CB230} This globule also hosts a single IRAS source (IRAS21169+6804), which is associated with the YSO found by Yun \\& Clemens (\\cite{yc94a}). The YSO is located at the apex of a cone-shaped nebulosity (see Fig.~\\ref{cb230h2}) and is at the centre of a CO outflow (Yun \\& Clemens~\\cite{yc92}; \\cite{yc94b}). The outflow is bipolar, but while the red component is quite compact, the blue lobe is more elongated, and extends northwards beyond the boundaries of the map made by Yun \\& Clemens (\\cite{yc94b}) (i.e. its length $>$3\\arcmin). Yun \\& Clemens (\\cite{yc94a}) found a second NIR object, \"possibly forming a binary system\" together with the YSO at the base of the outflow. In our observations we find this second NIR object to consist of two nuclei, which would make this a triple system (Fig.~\\ref{cb230h2}). The conical nebula seen in Fig.~\\ref{cb230h2} has a steep intensity-gradient towards the South, while towards the North the nebulous light fans out and decreases in intensity more gradually. The nebula opens up to the North, which is also the direction of the blue lobe of the outflow, the axis of which is North-South (Yun \\& Clemens~\\cite{yc94b}). The approaching component of the bipolar outflow may have cleared out a cavity; the more extended emission seen in the northern part of the nebula could then be scattered light from walls of this cavity (see Yun \\& Clemens~\\cite{yc94a}). Likewise, the diffuse emission associated with the embedded binary object $\\sim 9\\arcsec$ W of the main nebula is extended towards the SE (see Fig.~\\ref{cb230h2}). This might be the signature of an as yet undetected (or more likely: unresolved) outflow associated with the objects embedded herein. \\noindent Because we want to distinguish the (pure) line emission from the nebular emission, and because the spectral slope of the nebulosity differs from that of the field stars, we have determined the slope $x$ of the wavelength dependence of the emission ($\\propto \\lambda^x$) by using the integrated flux of the northern part of the nebula, excluding the region with the embedded star. As we shall see (e.g. Fig.~\\ref{feiih2closeup}) the consequence of this is that in the continuum-subtracted [FeII]-image the YSO is still visible. Its spectral index differs from that of the nebula (which is radiation scattered by the dust), which we have used to scale the H$_{\\rm cont}$-image; the YSO cannot therefore be correctly subtracted. \\noindent The final reduction, with careful scaling of the H$_{\\rm cont}$- and K$_{\\rm cont}$ images of the diffuse emission, reveals the pure line emission that remains after subtraction of the continuum, and is shown in Fig.~\\ref{feiih2closeup}: a jet is seen in [FeII], primarily defined by two knots along its length: a bright one at the tip (hereafter called k1) and a fainter one (k2) about half-way between k1 and the YSO. The knots are superimposed on a fainter, but clearly visible elongated emission feature. We stress that this detection is independent of potential inaccuracies in the continuum subtraction, as the knots are visible in the contours of the [FeII] image before continuum subtraction (cf. Fig.~\\ref{cb230h2}), but not at all in the (H$_{\\rm cont}$) continuum image alone. \\hfill\\break\\noindent The jet is oriented in the N-S direction, and lies at the base of the large-scale molecular outflow. This jet is also traced by H$_2$ emission (Fig.~\\ref{feiih2closeup}b), which is weaker and slightly less narrow than that of [FeII]. Moreover, the H$_2$ emission appears to be anti-correlated with [FeII] peaking next to, rather than on the secondary knot in [FeII], and we conclude that in H$_2$ we probably see the {\\sl walls} of the jet-channel. The detection of strong [FeII]- and weaker H$_2$ emission suggests the presence of fast, dissociative J-shocks. \\begin{figure*} \\resizebox{11cm}{!}{\\rotatebox{270}{ \\includegraphics{0931fg5a.eps}}} \\hspace{0.5cm} \\resizebox{6cm}{!}{\\rotatebox{270}{ \\includegraphics{0931fg5b.eps}}} \\caption[]{A close-up of the jets and knots in CB230, after subtraction of the continuum. {\\bf a}\\ [FeII]-emission, and {\\bf b}\\ the same but overlaid with H$_2$ contours (white). {\\bf c}\\ Comparison of cuts through centre of the knots in the [FeII]-jet at constant Dec-offset from the continuum-subtracted image. The profile for knot k1 is drawn as a thick histogram, that for k2 as a thin line. Most of the elevated emission plateau between about $-8\\arcsec\\ {\\rm and} +9\\arcsec$ is an artifact of the mosaicing together of the different frames constituting the final image. } \\label{feiih2closeup} \\end{figure*} \\noindent The total length of the jet, from its base at the location of the embedded YSO to the edge of knot k1 is about 9\\pas5, corresponding to $\\sim 0.02$~pc at the assumed distance of 450~pc (Launhardt \\& Henning~\\cite{launhen}). Profiles of the knots in the [FeII]-jet, along cuts made at constant declination offsets and passing through the location of peak intensity of the knots, are shown in Fig.~\\ref{feiih2closeup}c. Both features are resolved in the NIR image. The width (FWHM) of k1 is 1\\pas2\\, corresponding to $\\sim 540$~AU. In the profile of k2 one can also identify the two intensity enhancements to the East and West, seen in Fig.~\\ref{feiih2closeup}a, which deliniate the edges of the conical nebula. Both knots are superimposed on a broad ($\\sim 17\\arcsec \\approx 0.04$~pc) plateau of emission; this is an artifact due to imperfect mosaicing, rendering the middle section of the image (in which the nebula and the jet are located) more noisy than the edges. The absolute coordinates of the three stellar objects (A, B1, and B2; Fig.~\\ref{cb230h2}) and of the two knots (k1, k2; Fig.~\\ref{feiih2closeup}a) in the [FeII]-jet are given in table~\\ref{knotparams}, where we also list the integrated fluxes of the two [FeII]-knots and the entire H$_2$ jet. \\hfill\\break\\noindent In addition to the main jet, a second [FeII]-feature is seen, emanating from the bright nebulosity and oriented in a N-NW-direction; this may indicate the existence of a second (unresolved) outflow, or it is a shock at the surface of the cavity, and thus related to the main jet. \\begin{table} \\caption[] {List of coordinates and fluxes of the detected objects} \\label{knotparams} \\begin{tabular}{lcccc} \\hline \\multicolumn{1}{c}{Object} & \\multicolumn{1}{c}{$\\alpha_{\\rm 2000}$} & \\multicolumn{1}{c}{$\\delta_{\\rm 2000}$} & \\multicolumn{1}{c}{$F_{\\rm [FeII]}$} & \\multicolumn{1}{c}{$F_{\\rm H_2}$} \\\\ \\multicolumn{1}{c}{CB} & \\multicolumn{1}{c}{($^h$ $^m$ $^s$)} & \\multicolumn{1}{c}{($\\degr$ $\\arcmin$ $\\arcsec$)} & \\multicolumn{2}{c}{(erg cm$^{-2}$ s$^{-1}$)$^{\\dagger}$} \\\\ \\hline 3-k1 & 00 28 42.17 & +56 41 57.36 & $<$ 2.4$-$16 & 3.5$-$14 \\\\ 3-k2 & 00 28 42.12 & +56 41 37.17 & $<$ 2.4$-$16 & 1.4$-$14 \\\\ 3-k3 & 00 28 41.75 & +56 41 34.20 & $<$ 2.4$-$16 & 5.6$-$15 \\\\ 3-k4 & 00 28 42.12 & +56 41 28.56 & $<$ 2.4$-$16 & 2.1$-$14 \\\\ 205-A & 19 45 24.15 & +27 50 57.18 & -- & -- \\\\ 205-B & 19 45 23.99 & +27 50 59.25 & -- & -- \\\\ 230-A & 21 17 38.36 & +68 17 32.87 & -- & -- \\\\ 230-B1 & 21 17 40.09 & +68 17 32.15 & -- & -- \\\\ 230-B2 & 21 17 39.96 & +68 17 31.91 & -- & -- \\\\ 230-H$_2$ & -- & -- & -- & 1.1$-$14 \\\\ 230-k1 & 21 17 38.11 & +68 17 41.04 & 6.4$-$15 & -- \\\\ 230-k2 & 21 17 38.12 & +68 17 38.08 & 3.2$-$15 & -- \\\\ \\hline \\multicolumn{5}{l}{$\\dagger$ a$-$b means a$\\times 10^{-b}$} \\end{tabular} \\begin{center} \\end{center} \\end{table} \\subsection{CB188} \\begin{figure*} \\resizebox{15cm}{!}{\\rotatebox{270}{ \\includegraphics{0931fig6.eps}}} \\caption[]{Results of the TNG/NICS observations showing the H$_2$ emission, still including continuum emission, towards CB188 (left panel) and CB205 (right panel). The locations of the IRAS point sources are indicated by their uncertainty ellipses. The upper small panel on the far right shows a blow-up of the region around the IRAS source position in CB205. The lower small panel on the far right shows the alleged jet-like feature (Yun et al.~\\cite{yetal93}) being resolved into stars. } \\label{cb188205} \\end{figure*} In this globule, Launhardt \\& Henning (\\cite{launhen}) detected 1.3~mm continuum emission, in the form of an ``extended source with compact components'' at the location of the IRAS source (indicated by the ellipse in Fig.~\\ref{cb188205}). A YSO was detected at the same location by Yun \\& Clemens (\\cite{yc95}), while Yun \\& Clemens (\\cite{yc94b}) mapped a small ($\\leq 2\\arcmin$) outflow, of which the less extended blue lobe overlaps the red one. \\noindent Our continuum-subtracted H$_2$ and [FeII]-images do not show any line emission in the field down to the sensitivity limit. We also do not see the cometary diffuse emission around the possible NIR counterpart of the IRAS source that was found by Yun \\& Clemens (\\cite{yc94a}; this would be the bright star seen just outside the IRAS error ellipse in Fig.~\\ref{cb188205}), nor do we detect their source F, which they find embedded in this diffuse emission (see their Fig.~10b). It might be either a spurious detection or a variable object. \\subsection{CB205} This globule, also known as L810, has an extended NIR nebulosity at the location of the IRAS source, as seen in Fig.~\\ref{cb188205}. A NIR star cluster is associated with it; at least 10 stars can be seen within $10\\arcsec$ of the IRAS position (see the top-inset of Fig.~\\ref{cb188205}). \\hfill\\break\\noindent The major axis of the nebula is oriented in a N-S direction, which is also the orientation of the compact molecular outflow detected by Yun \\& Clemens (\\cite{yc94b}), of which the red and blue lobes show a significant overlap. This coincidence of orientations suggests a connection between the shape of the nebula and the presence of the outflow. However, no line emission has been found in our continuum-subtracted images, although we cannot rule out that some line emission knots may have been hidden in the more intense parts of the diffuse NIR nebulosity, near the IRAS point source location. \\noindent The nebula has been studied in the NIR (J,H,K) by Yun et al. (\\cite{yetal93}). The object they identified as the illuminator of the nebula, L810IRS, coincides with our star B in Fig.~\\ref{cb188205}. Yun et al. (\\cite{yetal93}) also report the detection of an elongated jet-like feature, located about $35\\arcsec$ from L810IRS, to the SW. This feature is also visible in all our images (both line and continuum) and is seen to be resolved in a coincidental alignment of stars embedded in diffuse emission (see the lower-inset in Fig.~\\ref{cb188205}, at $\\Delta\\alpha \\sim -20\\arcsec$ and $\\Delta\\delta \\sim -45\\arcsec$). No emission is left after continuum subtraction, further excluding a jet origin for it." }, "0403/astro-ph0403415_arXiv.txt": { "abstract": "{ We present a quantitative study of massive stars in the High Excitation Blob N81, a compact star forming region in the SMC. The stellar content was resolved by HST and STIS was used to obtain medium resolution spectra. The qualitative analysis of the stellar properties presented in Heydari-Malayeri et al.\\ (\\cite{papI}) is extended using non-LTE spherically extended atmosphere models including line-blanketing computed with the code CMFGEN (Hillier \\& Miller \\cite{hm98}), and the wind properties are investigated. The main results are the following: \\begin{itemize} \\item The SMC-N81 components are young ($\\sim$ 0--4 Myrs) O stars with effective temperatures compatible with medium to late subtypes and with luminosities lower than average Galactic O dwarfs, rendering them possible ZAMS candidates. \\item The winds are extremely weak: with values of the order 10$^{-8}$/10$^{-9}$ \\myr\\, the mass loss rates are lower than observed so far for Galactic dwarfs. Only the recent study of SMC stars by Bouret et al.\\ (\\cite{jc03}) show the same trend. The modified wind momenta ($\\dot{M}$ v$_{\\infty}$ $\\sqrt{R}$) are also 1 to 2 orders of magnitude lower than observed for Galactic stars. Both the mass loss rates and the modified wind momenta are lower than the predictions of the most recent hydrodynamical models. \\end{itemize} The accuracy of the UV based mass loss rate determination, relying in particular on the predicted ionisation fractions, are carefully examined. We find that $\\dot{M}$ could be underestimated by a factor of up to 10. Even in this unlikely case, the above conclusions remain valid \\textit{qualitatively}. The reasons for such weak winds are investigated with special emphasis on the modified wind momenta: \\begin{itemize} \\item There may be a break-down of the wind momentum - luminosity relation (WLR) for dwarf stars at low luminosity (log L/L$_{\\odot}$ $\\la$ 5.5). However, reasons for such a breakdown remain unknown. \\item The slope of the WLR may be steeper at low metallicity. This is predicted by the radiation driven wind theory, but the current hydrodynamical simulations do not show any change of the slope at SMC metallicity. Moreover, there are indications that some Galactic objects have wind momenta similar to those of the SMC stars. \\item Decoupling may take place in the atmosphere of the SMC-N81 stars, leading to multicomponent winds. However, various tests indicate that this is not likely to be the case. \\end{itemize} The origin of the weakness of the wind observed in the SMC-N81 stars remains unknown. We suggest that this weakness may be linked with the youth of these stars and represents possibly the onset of stellar winds in recently formed massive stars. ", "introduction": "\\label{s_intro} Massive stars play key roles in various astrophysical contexts all along their evolution: they ionise ultra-compact HII regions while still embedded in their parental molecular cloud; they create ionised cavities and shape the surrounding interstellar medium during the main fraction of their lifetime; they experience strong episodes of mass loss when they become Luminous Blue Variables and Wolf-Rayet stars, revealing their core and enriching the ISM in products of H and He burning; they end their life as supernovae, producing the heavy elements and releasing large amounts of mechanical energy. During all these phases, massive stars lose mass through winds driven by radiation pressure on metallic lines. This affects not only their evolution (e.g. Chiosi \\& Maeder \\cite{cm86}) but also the surrounding interstellar medium in which the release of mechanical energy can trigger instabilities leading to the collapse of molecular clouds and to star formation. Moreover, bubbles and superbubbles observed on galactic scales are powered by such mass ejections. Hence, various astrophysical fields require the knowledge of quantitative wind properties of massive stars. Several studies have been carried out in the last two decades to determine these properties. At solar metallicity, the observational determinations (e.g. Howarth \\& Prinja \\cite{hp89}, Puls et al.\\ \\cite{puls96}, Herrero et al.\\ \\cite{hpv00}) are on average in good agreement with the most recent hydrodynamical predictions based on the radiation driven wind theory (Vink et al.\\ \\cite{vink00}), both in terms of mass loss rate and of the modified wind momentum - luminosity relation (WLR, e.g. Puls et al.\\ \\cite{puls96}) which quantifies the strength of the wind. At non solar metallicities, we expect the wind properties to vary with $Z$ due to the modified radiative acceleration through metallic lines. In particular, the mass loss rate should be proportional to $Z^{r}$ (Abbott \\cite{abbott82}, Puls et al.\\ \\cite{psl00}) and the WLR should be shifted towards lower values and should have a steeper slope. The most recent theoretical results predict $r \\sim 0.8$ (Vink et al.\\ \\cite{vink01}) but no change in the slope of the WLR, at least for $Z > 10^{-3} Z_{\\odot}$ (Hoffmann et al.\\ \\cite{tadziu02}, Kudritzki \\cite{kud02}). Observational studies indicate a reduction of the mass loss rate and of the terminal velocity in the Magellanic Clouds, but given the small number of objects studied so far, the behaviour of the WLR at low metallicity is still poorly understood. Several groups are currently analysing stars in sub solar (Crowther et al.\\ \\cite{paul02}, Hillier et al.\\ \\cite{hil03}, Bouret et al.\\ \\cite{jc03}) and super solar (Najarro et al.\\ \\cite{paco} , Figer et al.\\ \\cite{figer}) regions for a better understanding of wind properties in different environments. The present work on SMC-N81 stars takes part in this effort. The SMC-N81 region belongs to the class of the ``High Excitation Blobs'' (HEB) first introduced by Heydari-Malayeri \\& Testor (\\cite{ht82}). These blobs are compact regions of star formation in the Magellanic Clouds (see Heydari-Malayeri \\cite{mhmiau} for a complete review). They have a typical radius of a few pc and display the features of star forming regions: HII cavities, turbulent structures, ionisation fronts and shocks. Recent HST observations (Heydari-Malayeri et al.\\ \\cite{pap0}) have revealed for the first time its stellar content, 4 of the brightest stars being grouped in the central 2 \\arcsec\\ wide region. Subsequently spectra of the main exciting stars have been obtained with STIS onboard HST. The qualitative analysis of these spectra, presented in Heydari-Malayeri et al.\\ (\\cite{papI}, hereafter paper I), have already revealed interesting properties. First, the stars have been identified as mid O dwarfs with surprisingly low luminosities compared to ``classical'' dwarfs. Second, the UV spectra have shown signatures of very weak winds, even weaker than those usually observed in the SMC. These characteristics have lead Heydari-Malayeri et al.\\ (\\cite{papI}) to propose that the SMC-N81 stars could belong to the class of Vz stars which are massive stars thought to lie very close to the ZAMS (Walborn \\& Parker \\cite{wp92}). As such the properties of these stars, showing unusually weak winds compared to other SMC O stars, seem already quite interesting. Furthermore the association of these objects with a compact star forming region, presumably indicative of a very young age, allows one also to obtain unique constraints on properties of very young massive stars shortly after their birth. In fact such observations appear crucial for a better understanding of the earliest evolutionary phases of massive stars and to constrain their formation process which is still under debate (they may form by accretion on a protostellar core -- Norberg \\& Maeder \\cite{nm}, Behrend \\& Maeder \\cite{raoul}-- or by collisions between low mass components in dense stellar clusters -- Bonnell et al.\\ \\cite{bonnell}). With such objectives in mind we have carried out a quantitative study of the UV spectra of the SMC-N81 stars. First results have been presented in Martins et al.\\ (\\cite{lanzarote}). In fact we are able to determine upper limits on the mass loss rates of four O stars in this region, which turn out to be surprisingly low (typically $\\mdot \\la$ a few $10^{-9}$ \\myr) compared to predictions of the radiation driven wind theory, even when taking metallicity effects into account. Although no precise physical explanation is found for this behaviour we strongly suggest that this behaviour is related to the very youth of these massive stars. The remainder of the paper is structured as follows. Section \\ref{observations} briefly summarises the observations and data reduction. Section \\ref{cmfgen} describes the main ingredients of the modeling. In Sect.\\ \\ref{interstellar} we explain how interstellar lines are taken into account. The main results are given in Sect.\\ \\ref{analysis} and discussed in Sect. \\ref{neb_ste_prop} (nebular and stellar properties) and \\ref{wind_properties} (wind properties). Finally, Sect.\\ \\ref{conclusion} summarises the main results. ", "conclusions": "\\label{conclusion} Based on UV spectral obtained with STIS/HST we have analysed the stellar and wind properties of the four main exciting stars of the High Excitation Blob SMC-N81 using extensive calculations of spherically expanding non-LTE line blanketed atmosphere models with the code CMFGEN. The main results are the following: \\begin{itemize} \\item[$\\diamond$] {The stellar properties (L, \\teff) indicate that the SMC-N81 components are young ($\\sim$ 0--4 Myrs old) O stars which shows, with perhaps the exception of star 1, a lower luminosity than ``normal'' Galactic O dwarfs. This, together with the closeness to the ZAMS for star 3 and 11, confirms the conclusion of paper I that they may belong to the Vz class (Walborn \\& Parker \\cite{wp92}). } \\item[$\\diamond$] {The UV spectra of the N81 stars show unusually weak stellar winds. The upper limits on mass loss rates are of the order a few $10^{-9}$ \\myr\\ which is low compared to 1) Galactic stars of the same luminosity and 2) the most recent predictions of $\\dot{M}$ as a function of stellar parameters and metallicity. Point 1) could be qualitatively understood due to the reduced metallicity of the SMC but point 2) indicates that this reduction is higher than expected.} Although the mass loss rates derived from the UV line analysis are potentially affected by uncertainties in the modeled ionisation fractions, various tests indicate that the above conclusions remain qualitatively valid. \\item[$\\diamond$] {Our objects show modified wind momenta ($M_{\\odot} v_{\\infty} R^{1/2}$) which are, for the same luminosity $L$, lower by typically two orders of magnitude compared to the ``normal'' O star samples. Similarly low wind momenta have also been found by Bouret et al.\\ (2003) for 3 SMC stars in NGC 346. The modified wind momentum - luminosity relation of all the SMC objects could be interpreted as showing a break-down at low luminosities or a different slope than the Galactic relation. The current sample of SMC stars may indeed indicate a steeper slope at least for giants and dwarfs, but the scatter is still too large to firmly establish this trend. However, the most recent hydrodynamical models (Vink et al.\\ \\cite{vink01}, Kudritzki \\cite{kud02}, Hoffmann et al.\\ \\cite{tadziu02}) do not predict such a change in the slope between solar and SMC metallicities. Furthermore we present the first indications that some Galactic objects have also low wind momenta comparable to the SMC dwarfs. This also tends to exclude explanations based uniquely on metallicity.} \\item[$\\diamond$] Possible explanations for a breakdown of the modified wind momentum - luminosity relation at low luminosities are discussed. Ionic decoupling appears unlikely according to various estimates. A failure of the CAK parameterisation in high density atmospheres, discussed by Owocki \\& Puls (\\cite{op99}), might be invoked to explain a lower acceleration in the transsonic region where the mass loss rate is set. Although the physical mechanism leading to such weak winds remains currently unknown, we speculate that the low mass loss rate is probably intrinsically related to the youth of the stars, possibly testifying of a phase of the ``onset'' of radiatively driven winds in young O stars shortly after their formation. \\end{itemize} Further studies of very young massive stars, Vz stars, and related objects with indications of weak winds will be of great interest to attempt to understand these puzzling wind properties and to provide interesting constraints on the development of stellar winds in the early phases of massive star evolution or possibly even on the final phases of their birth. \\appendix" }, "0403/astro-ph0403690_arXiv.txt": { "abstract": "We present three-dimensional simulations of viscous dissipation of AGN induced gas motions and waves in clusters of galaxies. These simulations are motivated by recent detections of ripples in the Perseus and Virgo clusters. Although the sound waves generated by buoyant bubbles decay with distance from the cluster center, we show that these waves can contribute substantially to offsetting the radiative cooling at distances significantly exceeding the bubble size. The energy flux of the waves declines more steeply with radius than the inverse-square law predicted by energy conservation, implying that dissipation plays an important role in tapping the wave energy. We show that such dispersing sound waves/weak shocks are detectable as ripples on unsharp-masked X-ray cluster maps, and point out that the interfaces between the intracluster medium and old bubbles are also clearly detectable in unsharp-masked X-ray maps. This opens up the possibility of detecting fossil bubbles that are difficult to detect in radio emission. This mode of heating is consistent with other observational constraints, such as the presence of cool rims around the bubbles and the absence of strong shocks. Thus, the mechanism offers a way of heating clusters in a spatially distributed and gentle fashion. We also discuss the energy transfer between the central AGN and the surrounding medium. In our numerical experiments, we find that roughly 65 per cent of the energy injected by the AGN is transferred to the intracluster medium and approximately 25 percent of the injected energy is dissipated by viscous effects and contributes to heating of the gas. The overall transfer of heat from the AGN to the gas is comparable to the radiative cooling losses. The simulations were performed with the FLASH adaptive mesh refinement code. ", "introduction": "The long-standing problem of cooling flow clusters of galaxies, in which the central cooling time is much shorter than the Hubble time, is how to prevent the intracluster medium (ICM) from collapsing catastrophically on a short timescale. The original idea for maintaining the overall cluster stability (Fabian 1994) was to postulate that a certain amount of gas decouples from the flow and does not contribute to the cooling of the remaining gas. This model would require up to 1000 $M_{\\odot}$ yr$^{-1}$ in mass deposition rates to guarantee cluster stability. This has been found to be inconsistent with recent {\\it Chandra} (e.g., McNamara et al. 2000, Blanton et al. 2001) and XMM-{\\it Newton} observations (e.g., Peterson et al. 2001, 2003; Tamura et al. 2001). {\\it Chandra} observations reveal a number of clusters with X-ray cavities/bubbles created by the central active galactic nuclei (AGN). It has been suggested by many authors that AGN feedback may play a crucial role in self-regulating cooling flows (e.g. Churazov et al. 2001, Ruszkowski \\& Begelman 2002, Brighenti \\& Mathews 2003). One of the main outstanding issues is how the AGN heating comes about in detail. In principle, strong shocks generated by AGN outbursts can dissipate in the ICM and heat the gas. However, imaging observations of cooling flow cores do not give evidence for this mode of heating. Recent {\\it Chandra} observations of two well-known clusters, the Perseus cluster (Fabian et al. 2003a,b) and the Virgo cluster (Forman et al. 2004), suggest that dissipation of sound waves and weak shocks could be an important source of gas heating --- an idea first proposed by Fabian et al. (2003a). Further support for the idea that viscosity may play an important role in the ICM comes from a recent study of density profiles in clusters (Hansen \\& Stadel 2003). Recently, a number of papers have described simulations of bubble-heated clusters (e.g., Churazov et al. 2001, Br\\\"{u}ggen et al. 2002, Br\\\"{u}ggen \\& Kaiser 2002, Br\\\"{u}ggen 2003, Quilis et al. 2001). Numerical simulations of viscous dissipation of AGN energy in ICM were previously considered by Ruszkowski et al. (2004) (Paper I) and Reynolds et al. (2004).\\\\ \\indent The main purpose of this paper is to extend our previous work on viscous heating of the ICM by waves to three dimensions. The results of a simulation of viscous dissipation in three dimensions could differ from our previous two-dimensional results given that the amplitudes of waves decrease faster with radius in three dimensions. This would directly affect the spatial distribution of the viscous dissipation rate. Apart from performing our simulations in three dimensions, we extend our previous analysis and that of Reynolds et al. (2004) to include aspects of heating that were previously neglected. First, we show that the conclusions drawn from our 2D simulations carry over to three dimensions. Second, our new results include (i) X-ray maps and unsharp-masked X-ray images, (ii) extend the discussion of spatial distribution of energy dissipation and overall heating rate, (iii) present details of the flow of heat between the bubbles and the intracluster medium and (iv) discuss the wave decay rates.\\\\ \\indent The outline of this paper is as follows. In the next section we describe the assumptions of the model. Section 3 presents and discusses our results, focusing on the spatial distribution of heating. In particular, we discuss the detectability of the ripples/sound waves, their decay rate with distance from the center/bubble surface, the energy transfer from the AGN to the kinetic and thermal energy of the ICM, and the fraction of the dissipated energy that heats the gas. The fourth section discusses the limitations of our model. The final section summarizes our findings. ", "conclusions": "To summarize, we have analyzed the energy deposition in the cluster due to rising bubbles, sound waves and weak shocks. This was motivated by the recent discovery of such waves in the Perseus cluster by Fabian et al. (2003a) and in the Virgo cluster by Forman et al. (2004). We found that the dissipated energy may be comparable to the cooling rate, thereby significantly affecting the cooling flow or even quenching it altogether. We showed that about 65 per cent of the energy injected by the central source can be transferred to the ICM. Approximately 25 per cent of the energy injected by the AGN can be converted to heat, assuming Spitzer viscosity. We discussed the wave decay rates and showed that a significant fraction of wave energy is deposited within the cooling radius. The computed decay rates are consistent with linear theory estimates of the damping length. The damped sound waves or weak shocks are still detectable in unsharp-masked X-ray images. Old bubbles become increasingly difficult to detect in the X-ray maps as the contrast between the rising bubbles and the surrounding gas diminishes. However, apart from sound waves and weak shocks in unsharp-masked X-ray maps, the interfaces between the intracluster medium and old bubbles are also clearly visible. This opens up the possibility of detecting fossil bubbles that are difficult to detect in radio emission." }, "0403/astro-ph0403373_arXiv.txt": { "abstract": "The progress of optical astronomy in post-apartheid South Africa is discussed. Particular emphasis is given to the socio-political climate which embraced the idea of a 10-m class telescope as a flagship project that would lead to widespread development in science, technology and education - not only in South Africa, but across the subcontinent. ", "introduction": "This account of optical astronomy in South Africa starts where Feast (2002, hereafter Paper I) left off, in 1994 with the first democratic elections and the start of a new era. The end of apartheid offered vastly increased opportunities for international collaborations among individuals and institutions, which the astronomy community was quick to take advantage of. Nevertheless, while the historical strength of astronomy laid a firm foundation for growth and success, the reasons why the discipline thrived and grew in the following decade were complex and essentially political. In the following I attempt a brief description of the policies and socio-political climate that have nurtured astronomy in South Africa, while acknowledging that no two individuals will see this in the same way and that any such account will be incomplete and probably idiosyncratic. The government-funded facilities available for optical and infrared astronomy in South Africa are described with an emphasis on the 10-m Southern African Large Telescope (SALT), due to be commissioned in early 2005. No attempt is made to describe detailed scientific projects or results, but it is worth noting that productivity remained high throughout the decade, with the South African Astronomical Observatory (SAAO) annual report, for example, recording well over 100 publications per year from its user community. A detailed account of research at the beginning of the period can be found in the compilation edited by Warner (1995). The challenge for the future is to redirect these efforts towards making effective use of SALT. Two significant transformations were initiated in South African astronomy during this decade: one involves the change to big telescope astronomy and is only just starting; the other is a broad ``Africanization\" of activities which is far from complete but well underway. Prior to 1994 the main interactions had been with the international community (excluding Africa), particularly that in the UK. Most of the local optical astronomers had been born, and many of them trained, outside of Africa. The post-1994 investment in astronomy came with the assumption that this would change - that astronomers would find ways to interact with and influence South African science and that a cohort of indigenous astronomers would be trained and nurtured. The challenge has been to do this in such a way that these young scientists are the peers of their international contemporaries and not merely tokens to fill quotas. Interestingly this challenge is being met through strengthening collaborations and partnerships, both nationally and internationally. ", "conclusions": "I want to finish by contextualizing the spending on astronomy in South Africa through comparing expenditure on SALT with that on the Hubble Space Telescope (HST). The cost of HST at launch in 1990 was \\$1.5 billion (or \\$1.59 billion in 1992 assuming 3 percent inflation) while that of SALT in 2002 terms is about \\$25 million. Using data from the World Bank\\footnote{www.worldbank.org/data/countrydata.html} we see that the USA GDP in 1992 was \\$6.262 trillion, while that of South Africa in 2002 was \\$104.2 billion. The cost of SALT compared to the GDP of South Africa, 0.025 percent, is almost identical to that of HST compared to the GDP of the USA. There can be no doubt that the South African astronomy community is extraordinarily fortunate in having access to this level of support. The politicization of science is often problematic for scientists, and much has been written about the inevitability of this process in connection with big science projects (e.g. Enard 2002). The above comparison with HST demonstrates that SALT is big science for Africa; the people of South Africa will have expectations of it that are comparable to those of Americans for HST. We South Africans will do well to follow the example of the Space Telescope Science Institute in ensuring that: \\begin{itemize} \\item observing time goes to the astronomers with the best projects, \\item those astronomers are fully empowered to do first rate science with SALT \\item and that the outcomes of their research are made accessible to the public and particularly to young Africans.\\end{itemize} It seems appropriate to give the last word to the President of South Africa, Thabo Mbeki. In opening the South African Pavilion at the 2000 World Expo in Hanover, Germany, President Mbeki said the following: \\begin{quotation}``{\\it Now, in the small town of Sutherland in the semi-desert Karoo region of our country, we are building a gigantic African eye through which we can view the universe. The construction of the single largest telescope in the southern hemisphere, SALT - as it is called - will mean that in this humble home of our earliest humans, we are also building a vast gateway through which we can observe our earliest stars, learn about the formation of our galaxy and the lives of other worlds so as to give us insights into our future. We are proud that SALT will not only enable South African scientists to undertake important research, but also provide significant opportunities for international collaboration and scientific partnerships with the rest of the world.}\"\\end{quotation}" }, "0403/astro-ph0403145_arXiv.txt": { "abstract": "The center of M83, a barred starburst galaxy with a double nucleus, has been observed in the CO($J$=2--1) and CO($J$=3--2) lines with the Submillimeter Array. The molecular gas shows a distribution and kinematics typical for barred galaxies at $\\sim$kpc radii, but reveals unusual kinematics around the double nucleus in the central $\\sim$300 pc. Our CO velocity data show that the visible nucleus in M83 is at least 3\\arcsec\\ (65 pc) away from the galaxy's dynamical center, which most likely coincides with the center of symmetry previously determined in $K$ band and is suggested to host another nucleus. We discovered high-velocity molecular gas associated with the visible off-center nucleus, and also found a steep velocity gradient across it. We attribute these features to a gas disk rotating around the off-center nucleus, which may be the remnant of a small galaxy accreted by M83. The dynamical mass of this component is estimated to be $3\\times 10^8 \\Msol$ within a radius of 40 pc. The dynamical perturbation from the off-center nucleus may have played a key role in shaping the lopsided starburst. ", "introduction": "M83 (NGC 5236) is a nearby face-on barred spiral galaxy with a double nucleus and a nuclear starburst ($D=4.5$ Mpc, $1\\arcsec = 22$ pc; Thim et al. 2003). The galaxy is exceptionally complex in the central 300 pc, despite the symmetric appearance of its bar and spiral arms. In $K$ band, the brightest point, called the `visible nucleus', is offset by 3\\farcs4 (75 pc) from the centroid (i.e., center of symmetry) determined from the isophotes at radii of 0.3 -- 0.7 kpc \\citep[hereafter TTG]{Thatte00}. The centroid suffers from large extinction, and does not show a peak in $K$ band. However, the stellar velocity dispersion peaks both at the visible nucleus and near the isophotal centroid, and suggests a mass of $\\sim$10$^{7}$\\Msol\\ for each. This suggests a double nucleus (TTG). The visible nucleus has a hard power-law spectrum in X-ray due either to an accreting supermassive black hole or to X-ray binaries \\citep{Soria02}. \\citet{Elmegreen98} found a double circumnuclear ring in their $J-K$ color index map. The `inner' and `outer' rings of 190 pc and 60 pc radius are not concentric. The galaxy hosts a circumnuclear starburst mainly in the `starburst arc' of $\\sim$250 pc length that lies between the two rings \\citep{Gallais91, Harris01}. The galactic center has a large amount of molecular gas, estimated to be $10^{7.5\\mbox{--}8.5}\\Msol$ in $r\\leq300$ pc \\citep{Handa90,Israel01}. The proximity, face-on configuration, abundant molecular gas, starburst, and double nucleus make M83 one of the most interesting targets to study spiral galaxies and their nuclear activity through molecular gas. We observed M83 as a part of the early science program of the Submillimeter Array (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. }. The new telescope in Hawaii can easily observe the galaxy at $\\delta=-30\\degr$. In this {\\it Letter}, we first describe the observations and the bar-driven gas dynamics in the central 2 kpc, and then show that the visible nucleus is indeed offset from the dynamical center of M83 and has a velocity feature indicative of its own gas disk. We discuss the origin of the double nucleus, and its relation to the starburst activity. ", "conclusions": "" }, "0403/astro-ph0403235_arXiv.txt": { "abstract": "We present the analysis of 3 hrs of a rapid time series of precise stellar radial velocity (RV) measurements ($\\sigma$ = 4.5 m\\,s$^{-1}$) of the cool Ap star $\\beta$ CrB. The integrated RV measurements spanning the wavelength interval 5000-6000\\,{\\AA} show significant variations (false alarm probability = 10$^{-5}$) with a period of 16.21 min ($\\nu$ = 1028.17 $\\mu$Hz) and an amplitude of 3.54 $\\pm$ 0.56 m\\,s$^{-1}$. The RV measured over a much narrower wavelength interval reveals one spectral feature at $\\lambda$6272.0 {\\AA} pulsating with the same 16.21 min period and an amplitude of 138 $\\pm$ 23 m\\,s$^{-1}$. These observations establish $\\beta$ CrB to be a low-amplitude rapidly oscillating Ap star. ", "introduction": "Kurtz (1989) first suggested that the cool magnetic Ap star $\\beta$ CrB was a prime candidate to be a rapidly oscillating Ap (roAp) star since it had stellar properties similar to several known oscillating Ap stars (HR 1217, 33 Lib, and $\\gamma$ Equ). The issue of whether $\\beta$ CrB is a pulsating star has important implications for the excitation mechanism in roAp stars as well as the possible existence of an instability strip for roAp stars. Several investigators have searched for pulsational variations in $\\beta$ CrB using photometric measurements with null results (Heller \\& Kramer 1988; Kreidl 1991). Recently, Kochukhov et al. (2002, hereafter K02) presented radial velocity (RV) measurements for $\\beta$ CrB spanning the wavelength interval 6105 -- 6190 {\\AA}. Measurements of most lines were constant to a level of 20--30 m\\,s$^{-1}$; however, intriguing evidence for RV variations were found for one spectral feature, Fe I $\\lambda$6165.4 {\\AA}. This feature had an amplitude of 71 $\\pm$ 11 m\\,s$^{-1}$ and a period of 11.5 $\\pm$ 0.5 min. Before we can place $\\beta$ CrB in the family of roAp stars the K02 result must be confirmed since the false alarm probability for the detection was rather high (= 0.016). Here we present our own precise radial velocity measurements for $\\beta$ CrB and show that it is indeed a rapidly oscillating Ap star, but not with the period found by K02. ", "conclusions": "Our precise RV measurements for $\\beta$ CrB establish that this star is with high probability a rapidly oscillating Ap star with an amplitude of 3.54 $\\pm$ 0.56 m\\,s$^{-1}$ and a period of 16.2 $\\pm$ 0.7 min. Although additional observations are needed to confirm our result, we believe that the signal we have detected is real for a number of reasons: \\begin{itemize} \\item {\\it We derive a very low false alarm probability.} The probability that this is a false signal due to noise is $\\approx$ 10$^{-5}$ as determined through the bootstrap randomization procedure. This is a more rigorous way of determining the FAP than strictly using the Lomb-Scargle power. This result represents the combined mean velocity of all spectral lines spanning the wavelength interval 5000--6000\\,{\\AA}. Any systematic or instrumental error would have to affect all wavelength chunks used in the analysis in the same way which seems unlikely. \\item {\\it The variations are not due to a few outliers.} We think it is unlikely that a few outliers are causing our signal. (We define an outlier as deviating significantly from the mean RV value and not from the best fit sine wave through the data.) If only a few outliers are driving the power in the periodogram then this would be evident in a high false alarm probability. For a real periodic signal in the presence of noise, the more data one accumulates the more significant the detection becomes, even with the presence of a few occasional outliers. Thus the power in a Lomb-Scargle periodogram should increase with increasing number of data points. This is demonstrated in Figure~\\ref{powerinc} which shows the Lomb-Scargle power at $\\nu$ = 1028.17 $\\mu$Hz (circles) as a function of the number of data points used in the periodogram. For comparison we generated a synthetic signal consisting of a sine wave with the same period and amplitude as found in our data and sampled in the same manner as the real data. Random noise was added with the same rms scatter as our measurements. The crosses in Figure~\\ref{powerinc} show that the L-S power of the fake data (with a signal present) increases with increasing number of data points in the same way. An iodine absorption cell was used to perform our RV measurements because it is designed to eliminate, or at least minimize instrumental effects. However, one could argue that there might be residual instrumental shifts not corrected by the reduction process, or that our instrumental profile modelling somehow introduces a false signal into the data. We do not believe this is the case. We investigated whether periodic instrumental shifts were present in our data using the wavelength solution calculated from the iodine absorption lines. A Fourier analysis of these shifts showed the strongest period at 176 minutes, and smaller variations with periods of 20 and 24 minutes. All periods were significantly different from the one found in our RV data. Furthermore, we also performed a Fourier analysis of all IP parameters used in our RV determination. We found no variations at the period coinciding with the one found in the RV data. This demonstrates that using the iodine cell technique does an excellent job of excluding instrumental variations. \\item {\\it The same signal is found in the $\\lambda$6272 {\\AA} feature.} The spectral orders used to determine the integrated RV did not include the order covering the $\\lambda$6272 {\\AA} feature. Thus this spectral line is not contributing to the signal found in our integrated RV measurements. However, a separate analysis of this spectral chunk did find the same 16.21 min period and with the same statistical significance as with the integrated RV measurements. Although this is not an independent confirmation (we are using the same data set, but different parts of the spectrum), this result and the above arguments argue strongly that the 16.21 min oscillations in $\\beta$ CrB are real. \\end{itemize} \\begin{figure} \\epsfxsize=8.5truecm \\epsffile{figure5.eps} \\caption{The Lomb-Scargle power at $\\nu$ = 1028.17 $\\mu$Hz as a function of number of data points used in the periodogram using real (circles) and fake data (crosses). The fake data was generated using a sine wave with the same amplitude (3.56 m\\,s$^{-1}$) and period (= 16.21 min) as found in our data and sampled in the same manner. Random noise with the same scatter ($\\sigma$ = 4.5 m\\,s$^{-1}$) as our data was also added to the fake signal. } \\label{powerinc} \\end{figure} Our RV measurements fail to confirm the results of K02 on the Fe I $\\lambda$6165\\,{\\AA} feature. The wavelength chunk containing this spectral line had an rms scatter of 130 m\\,s$^{-1}$ (due to the weak iodine absorption lines in this wavelength region). However, we had a factor of 3 more measurements than K02. Monte Carlo simulations indicate that we would\thave detected an 11.5 min period with an amplitude of 70 m\\,s$^{-1}$ and a FAP of 10$^{-4}$. An amplitude of 50 m\\,s$^{-1}$ would have been detected with a FAP = 0.01. Thus we would have detected any variations in the Fe I $\\lambda$6165\\,{\\AA} line if they were present at the same amplitude as reported by K02. We believe that the result of K02 on $\\beta$ CrB is spurious. The FAP is much too high (FAP = 0.016) and our measurements have failed to find the presence of an 11.5 min period in either the integrated RV measurements, or in the analysis of the narrow wavelength chunks. More troubling is that the K02 period of 11.5 $\\pm$0.5 min found in $\\beta$\\,CrB is very close to the values of periods ($\\approx$ 11.7 min) found in several spectral lines in the roAp star 10 Aql using the same instrument. The most likely explanation for the RV signal found in $\\beta$ CrB by K02 is that it is due to short term uncorrected instrumental variations in the Gecko spectrograph used for the measurements. Circumstantial evidence for this comes from Matthews \\& Scott (1995) who reported an 11.1 min period in the RV variations in $\\gamma$\\,Equ. This period is different from known pulsation modes and has never been confirmed by subsequent RV measurements from other investigations. Although the authors used a superimposed mercury emission line from an arc lamp to eliminate instrumental shifts, the 11.1 min period could still be a residual instrumental effect due to the stellar light a calibration source having a slightly different optical path (not the case when using the iodine absorption cell). The RV measurements of K02, on the other hand, were made {\\it without} a simultaneous wavelength calibration (either I$_2$ absorption cell or simultaneous Th-Ar calibration), rather the wavelength calibration used a Th-Ar exposure taken before and after the time series used for the RV measurements. This cannot correct for any instrumental variations on time scales significantly shorter than the time between the two calibration measurements. Because K02 have failed to exclude an instrumental origin with a period of $\\approx$ 11.5 min for the RV variations in $\\beta$ CrB their results on this star (and possibly 10 Aql) should be considered suspect. An investigation of the short-term instrumental shifts of the Gecko spectrograph would be useful. Our RV measurements for $\\beta$ CrB seem to establish that this star is an roAp star with the lowest RV amplitude (3.5 m\\,s$^{-1})$ and one of the longest periods (16.2 min). $\\beta$ CrB is also unique among roAp stars in that it shows high amplitude pulsational in only one spectral feature, the $\\lambda$6271.9\\,{\\AA}. This showed significant variations (FAP $=$ 1.5$\\times$10$^{-5} $) with the same 16.2 min period and an amplitude of 138 $\\pm$ 23 m\\,s$^{-1}$. We tentatively identify this feature as a blend of Ce\\,II and Cr\\,II lines. Why $\\beta$ CrB has such a low RV amplitude, almost no high amplitude spectral lines, and a longer period mode compared to other roAp stars with similar spectral properties should provide clues as to the origin of the roAp phenomenon. We are currently planning observations of this star over a full rotation period not only to confirm our detection, but to search for rotationally modulated amplitude variations of the pulsations." }, "0403/astro-ph0403529_arXiv.txt": { "abstract": "Resonant active-to-active ($\\nu_a \\rightarrow \\nu_a$), as well as active-to-sterile ($\\nu_a \\rightarrow \\nu_s$) neutrino ($\\nu$) oscillations can take place during the core bounce of a supernova collapse. Besides, over this phase, weak magnetism increases antineutrino ($\\bar{\\nu}$) mean free paths, and thus its luminosity. Because the oscillation feeds mass-energy into the target $\\nu$ species, the large mass-squared difference between species ($\\nu_a \\rightarrow \\nu_s$) implies a huge amount of energy to be given off as gravitational waves ($L_{\\textrm{GWs}} \\sim 10^{49}$~erg s$^{-1}$), due to anisotropic but coherent $\\nu$ flow over the oscillation length. This asymmetric $\\nu$-flux is driven by both the spin-magnetic and the {\\it universal spin-rotation} coupling. The novel contribution of this paper stems from 1) the new computation of the anisotropy parameter $\\alpha \\sim 0.1-0.01$, and 2) the use of the tight constraints from neutrino experiments as SNO and KamLAND, and the cosmic probe WMAP, to compute the gravitational-wave emission during neutrino oscillations in supernovae core collapse and bounce. We show that the mass of the sterile neutrino $\\nu_s$ that can be resonantly produced during the flavor conversions makes it a good candidate for dark matter as suggested by Fuller et {\\it al.} (2003). The new spacetime strain thus estimated is still several orders of magnitude larger than those from $\\nu$ difussion (convection and cooling) or quadrupole moments of neutron star matter. This new feature turns these bursts the more promissing supernova gravitational-wave signal that may be detected by observatories as LIGO, VIRGO, etc., for distances far out to the VIRGO cluster of galaxies. \\vskip 0.8 truecm ", "introduction": "{\\it Supernovae neutrinos and gravity waves.---} That outflowing neutrinos ($\\nu$s) from a supernova (SN) generate gravitational waves (GWs) was firstly pointed out by Epstein (1978). However, over the first $\\sim 10$ milliseconds (ms) (Mayle, Wilson \\& Schramm 1987; Walker \\& Schramm 1987) after the SN core bounce the central density gets so high that no radiation nor even $\\nu$s can escape, they are thus frozen-in and strongly coupled to the neutron matter ($N^0$) as described by the Lagrangean (see Kusenko \\& Postma 2002 for this dynamics) \\be L^{int}_{N^0\\leftrightarrow \\nu} = \\frac{G_F}{\\sqrt{2}} \\left[ \\bar{N}^0 \\gamma_\\mu(1-\\gamma_5)N^0 \\right] \\left\\{ \\bar{\\psi} \\gamma^\\mu(1-\\gamma_5)\\psi \\right\\}\\; ,\\label{interact} \\ee with the $\\nu$ field $(\\psi)$ satisfying the time-dependent Dirac equation \\be \\left[i\\gamma^0 \\partial_0 + i\\gamma^\\alpha \\partial_\\alpha + \\rho(t) v_\\beta \\gamma^\\beta \\left(\\frac{1-\\gamma_5}{2}\\right) - m_\\nu\\right] \\psi = 0. \\label{dirac-mass} \\ee At this phase the whole proto-neutron star (PNS) dynamics is dominated by gravity alone, and can be appropriately described by the general relativistic Oppenheimer-Volkoff equation for both the $N^0$ + $\\nu$ fluid (see Mosquera Cuesta 2002). As discussed by Mayle, Wilson \\& Schramm (1987); and Walker \\& Schramm (1987), it is over this early transient that most $\\nu$ flavor conversions are expected to resonantly take place and consequently the super strong GWs burst from the oscillation process to be released. GWs from this decoupling has been suggested to likely be the ultimate process responsible for the neat kick given to a nascent pulsar during the SN collapse (Mosquera Cuesta 2000; 2002). The contention of this \\textit{paper} is a) to pave, in the framework of general relativity (GR), the pathway to this fundamental astrophysical process of generation of GWs from $\\nu$ oscillations in a PNS. b) to demonstrate, by taking into account experimental and observational constraints, that $\\nu$ oscillations during SN core bounce do produce GWs of the sort predicted by Einstein's GR theory, and more crucial yet, c) to stress that these bursts are the more likely SN GWs-signals to be detected by interferometric observatories as LIGO, VIRGO, GEO-600, etc. We speculate that such a signal perhaps might have been detected during the SN1987a event, despite the low sensitivity of the detectors at the time. Some claims in this direction were presented by Aglietta, Amaldi, Pizzella, et al.\\cite{amaldi89}, and related papers. ", "conclusions": "One can see that if $\\nu$ flavor conversions indeed take place during SN core bounce inasmuch as they take place in our Sun and Earth (Smirnov 2002), then GWs should be released during the transition. The GWs signal from the process is expected to irradiate much more energy than current mechanisms figured out to drive the NS dynamics at birth do. A luminosity this large (Eq.(\\ref{GWs-luminosity})) would turn these bursts the strongest GWs signal to be detected from any SN that may come to occur, futurely, on distances up to the VIRGO cluster, $R \\sim [10-20]$ Mpc. It is stressed that this signal will still be the stronger one from a given SN, even in the worst case in which the probability of $\\nu$ conversion is three orders of magnitude smaller then the estimated in the present paper. In proviso, we argue that a GWs signal that strong could have been detected during SN1987a from the Tarantula Nebula in the Milky Way's satellite galaxy Large Magellanic Cloud, despite of the low sensitivity of the detectors at the epoch. In such a case, the GWs burst must have been correlated in time with the earliest arriving neutrino burst constituted of some active species given off during the very early oscillation transient where some $\\nu_e$s went into $\\nu_{\\mu}$s, $\\nu_\\tau$s or $\\nu_s$s. Thenceforth, it could be of worth to reanalyze the data collected for from that event taking careful follow up of their arrival times, if appropriate timing was available at that moment." }, "0403/astro-ph0403325_arXiv.txt": { "abstract": "The Sloan Digital Sky Survey has validated and made publicly available its Second Data Release. This data release consists of 3324 square degrees of five-band ($u\\,g\\,r\\,i\\,z$) imaging data with photometry for over 88 million unique objects, 367,360 spectra of galaxies, quasars, stars and calibrating blank sky patches selected over 2627 degrees of this area, and tables of measured parameters from these data. The imaging data reach a depth of $r \\approx 22.2$ (95\\% completeness limit for point sources) and are photometrically and astrometrically calibrated to 2\\% rms and 100 milli-arcsec rms per coordinate, respectively. The imaging data have all been processed through a new version of the SDSS imaging pipeline, in which the most important improvement since the last data release is fixing an error in the model fits to each object. The result is that model magnitudes are now a good proxy for point spread function (PSF) magnitudes for point sources, and Petrosian magnitudes for extended sources. The spectroscopy extends from 3800\\AA\\ to 9200\\AA\\ at a resolution of 2000. The spectroscopic software now repairs a systematic error in the radial velocities of certain types of stars, and has substantially improved spectrophotometry. All data included in the SDSS Early Data Release and First Data Release are reprocessed with the improved pipelines, and included in the Second Data Release. Further characteristics of the data are described, as are the data products themselves and the tools for accessing them. ", "introduction": "The Sloan Digital Sky Survey (SDSS; York \\etal\\ 2000) is an imaging and spectroscopic survey of the high Galactic latitude sky visible from the Northern hemisphere. The principal survey goals are to measure the large-scale distribution of galaxies and quasars and to produce an imaging and spectroscopic legacy for the astronomical community. The SDSS data have been used in well over 200 refereed papers to date on subjects ranging from the colors of asteroids (Ivezi\\'c \\etal\\ 2002) to magnetic white dwarfs (Schmidt \\etal\\ 2003) to structures in the Galactic halo (Newberg \\etal\\ 2003) to the star-formation history of galaxies (Kauffmann \\etal\\ 2003) to Type II quasars (Zakamska \\etal\\ 2003) to the large-scale distribution of galaxies (Pope \\etal\\ 2004; Tegmark \\etal\\ 2004). The survey uses a dedicated 2.5m telescope with a three-degree field of view at Apache Point Observatory, New Mexico. A 120 mega-pixel camera (Gunn \\etal\\ 1998) images in five broad bands ($u,g,r,i$ and $z$; Fukugita \\etal\\ 1996; Stoughton \\etal\\ 2002) on clear moonless nights of good seeing. These data are photometrically calibrated using an auxiliary 20-inch telescope with a $40^\\prime \\times 40^\\prime$ imager, which determines the photometricity of each night (Hogg \\etal\\ 2001), and measures the extinction and photometric zeropoint using a network of % standard stars (Smith \\etal\\ 2002). The imaging data are processed through a series of pipelines that locate and measure the properties of all detected objects (Lupton \\etal\\ 2001) and carry out photometric and astrometric calibration (Pier \\etal\\ 2003). From the resulting catalogs of objects, complete catalogs of galaxies (Eisenstein \\etal\\ 2001; Strauss \\etal\\ 2002) and quasar candidates (Richards \\etal\\ 2002) are selected for spectroscopic followup, and are assigned to spectroscopic tiles of diameter 3 degrees (Blanton \\etal\\ 2003). Spectroscopy is performed on nights with moonlight, mild cloud cover, and/or poor seeing using a pair of double spectrographs with coverage from 3800--9200\\AA, and resolution $\\lambda / \\Delta \\lambda$ of roughly 2000. A plug plate for each tile holds 640 optical fibers of $3''$ entrance aperture which feed the spectrographs, together with eleven coherent fiber bundles to image guide stars. Because of the diameter of the cladding holding the optical fibers, spectroscopy cannot be carried out for objects separated by less than $55''$ on a given plate. ", "conclusions": "" }, "0403/astro-ph0403439_arXiv.txt": { "abstract": "We present UKIRT UIST spectra of Sakurai's Object (=V4334~Sgr) showing CO fundamental band absorption features around 4.7~$\\mu$m. The line-centres are at heliocentric radial velocity of $-$170$\\pm$30~km~s$^{-1}$. The number and relative strengths of the lines indicate a CO gas temperature of $400\\pm100$~K and CO column density of 7$^{+3}_{-2}\\times10^{17}$~cm$^{-2}$. The gas was moving away from the central star at an average speed of $\\sim$290$\\pm$30~km~s$^{-1}$ in 2003~September. The lines appeared sometime between mid 1999 (well after the opaque dust shell formed) and mid 2000 and may have been somewhat more blue--shifted initially than they are now. The observed CO velocity and temperature indicate the continued presence of a fast wind in the object, previously seen in the He~{\\sc i} 1.083~$\\mu$m line beginning just prior to massive dust formation, and more recently in atomic and ionized lines. The dust continuum is consistent with a temperature of 350$\\pm$30~K, indicating continued cooling of the shell. The similar CO temperature suggests that the bulk of the CO absorption occurs just outside of the dust continuum surface. ", "introduction": "\\label{sec-intro} Stellar evolution theory accounts reasonably well for the effects of thermal pulses on the development of stars on the Asymptotic Giant Branch. Less well constrained are the effects of a late thermal--pulse (LTP, once the star has shed its envelope) or very--late thermal--pulse (VLTP), once the star has begun to descend the white dwarf cooling track). However, indications are that perhaps 10 to 20\\%\\ of low-- and intermediate--mass stars undergo LTP or VLTP. It appears that the subsequent ``born--again'' evolution across the HR diagram is short lived (perhaps a few centuries, but perhaps as short as a few decades; see \\citealt{Iben83,Iben96, Lawlor03}), and consequently very few are observable at any one time. Sakurai's Object (= V4334~Sgr) is probably the first example of a VLTP observable with non--optical instruments in the immediate post--flash epoch. Sakurai's~Object was first identified as possibly undergoing a helium--shell--flash on 1996~February~23 \\citep{Nakano96, Benetti96}, but it is clear that it started to increase in brightness in mid--1994 \\citep{Takemizawa97}. In 1995 we began an infrared (IR) spectroscopic monitoring programme using the United Kingdom Infrared Telescope (UKIRT) and the cooled grating spectrometer CGS4 \\citep{Mountain90}; more recently we have used the UKIRT Imaging Spectrometer UIST \\citep{Ramsay-Howat00}. Observations are carried out throughout the star's observable period each year and the latest data, taken in 2003~September, are presented here. ", "conclusions": "\\label{sec-conclusion} A recent spectrum of Sakurai's Object has clearly revealed the presence of highly blue--shifted absorption lines of the fundamental band of CO. The CO was first noted in a spectrum from 2003~September~8. Re--examination of previous lower resolution spectra shows that the CO features must have arisen between 1999~May~4 and 2000~April~17. The blue--shift of the lines is consistent with gas moving away from the central star at $\\sim$300~km~s$^{-1}$. The number and relative strengths of the lines suggest a CO gas temperature of $400\\pm100$~K and CO column density of 7$^{+3}_{-2}\\times10^{17}$~cm$^{-2}$ (the lower value corresponding to the higher temperature), and linewidths no greater than 25~km~s$^{-1}$ (FWHM). We rule out interstellar CO as the source of the absorption lines, but the circumstellar CO could be formed shortly before discovery or existed for some time prior to emerging from the dusty shroud. % We note that the CO velocity is additional evidence for the existence of a fast wind in this object, although we cannot make a direct connection with changes in the central star. Taking the 1--5~$\\mu$m spectrum overall we find that the dust continues to cool, and the temperatures of the CO--bearing and dust--bearing materials are similar. Hence the two components presumably lie at a similar distance from the central star. As the CO is moving more rapidly than the dust continuum surface, this suggests an ongoing wind replenishing the CO in the region of the dust." }, "0403/astro-ph0403263_arXiv.txt": { "abstract": "Aluminum and other metal abundances were determined in 21 red giants in the globular clusters NGC~6752 and M80 as part of a larger study to determine whether the aluminum distribution on the red giant branch is related to the second parameter effect that causes clusters of similar metallicity to display different horizontal branch morphologies. The observations were obtained of the Al~I lines near 6700 {\\AA} with the CTIO Blanco 4-m telescope and Hydra multi-object spectrograph. The spectra have a resolving power of 18000 or 9400, with typical S/N ratios of 100-200. Mean [Fe/H] values obtained from the spectra are --1.58 for NGC~6752 and --1.73 for M80; this represents the first spectroscopic iron abundance determination for M80. Both NGC~6752 and M80 display a spread in aluminum abundance, with mean [Al/Fe] ratios of +0.51 and +0.37, respectively. No trend in the variation of the mean Al abundance with position on the giant branch is discernible in either cluster with our small sample. ", "introduction": "\\label{sec:intro} Over 20 years have passed since \\citet{NCFD81} first showed aluminum abundance variations on the red-giant branch (RGB) of the globular cluster NGC~6752, and yet the nature of these inhomogeneities remains a mystery. Observations since then continue to show aluminum (and sodium and magnesium) variations in other clusters \\citep[see, e.g.][]{CD81,WLO87,SHFS87,DSS92, ND95,PSKL96,ZWB96,S96a,Kraft97,SKS97,K98,Ivans1999,CN2000,Ivans2001,RC2001,GratAl2001, GBNF2002}. Although variations in carbon and nitrogen were previously known, these could be described through a simple mixing mechanism proposed by \\citet{SM79}, where rotationally induced meridional circulation currents could carry nuclearly processed materials such as C, N, and O from around the hydrogen-burning shell (H shell) of a red giant across the radiative zone to the outer convective envelope. Heavier elements such as Mg, Na, and Al weren't thought to be processed around the H shell, and any variations in them were taken as evidence that some kind of primordial pollution affected the surface abundances \\citep[see, e.g.,][]{CD81}. Continued work on key nuclear reaction rates \\citep{Champagne98,DD90,Iliadis90, CBS93,Blackmon95,Iliadis96}, however, suggested that these elements could be processed around the H shell under the same conditions that the CN and ON nuclear cycles operated. Using these results, as well as the widely accepted rates of \\citet{CF88}, separate groups showed that it is possible to account qualitatively for the observed variations that showed Na, Al, and N anticorrelated with C, O, and, in some cases, Mg \\citep{LHS93, CSB96, DW96, CSB98, DDNW98}. The challenge has always been describing the results {\\em quantitatively}; in particular, producing [Al/Fe] as high as 1.5 dex without overproducing [Na/Fe], and depleting $^{24}$Mg to the observed levels in M13 \\citep{S96b}, all while remaining within the acceptable proton-capture rates \\citep{LHZ97,NACRE,Powell99,UNC}. Despite any latitude afforded by the reaction rate uncertainties \\citep[see, e.g.,][]{CSB98}, the mixing theory still relies on non-solar abundance ratios in the star prior to mixing \\citep{DDNW98,CN2000}, thus, at least partly relying on primordial influences. Theories other than meridional circulation have been put forth \\citep{LHZ97,FAK99, AFK2001,DW2001}, but a detailed discussion of each is beyond the intent of this paper. Regardless of the physics behind any mixing mechanism, the question still remains: are the Al (and Na, Mg) variations caused by mixing, primordial sources, or a combination of both? We might be able to answer this question by looking at another long outstanding problem in globular cluster astronomy, namely, the second-parameter effect. First pointed out by \\citet{SW67} and \\citet{vdB67}, the second-parameter effect refers to the phenomenon where the horizontal-branches (HB) of two clusters with similar metallicity (the first parameter) have markedly different color distributions. Possible second parameters that have been investigated include age, initial helium abundance, CNO abundance, and mass loss \\citep[see, e.g.,][]{Faulkner66,Renzini77,Chaboyer98}, among others, but none is applicable to all clusters. One recent suggestion hypothesizes that if deep mixing (i.e., mixing that penetrates the H shell) occurs, then helium will be brought to the surface affecting the HB morphology by making mixed stars both bluer and brighter (Sweigart 1997a,b). Unfortunately, helium cannot be measured in RGB stars because of their low surface temperatures, and helium settling on the HB precludes an accurate measurement in the hotter stars. However, models by \\citet{CSB98}, which use standard (albeit, uncertain) reaction rates, show that aluminum can be produced only in the H shell of giants within the last magnitude of the RGB, implying that an increase in helium in the envelope must also be accompanied by an increase in aluminum. From this we can postulate that, if [Al/Fe] variations are produced internally and not primordially, Al could be a good surrogate to measure He mixing on the bright RGB, and if He mixing is indeed the second parameter, a relationship should exist between the ratio of Al-strong to Al-normal stars (the Al ratio) and the ratio of blue to red HB stars (the HB ratio). It is the {\\em distribution} of abundances as a function of magnitude that is critical in determining the whether this correlation exists and what its cause might be. This current work attempts to provide some fresh data on clusters that have been historically under-studied. Both M80 (NGC~6093) and NGC~6752 possess blue HBs relative to other clusters at similar metallicity. For example, \\citet{Ferr98} compared Hubble $U,V$ M80 photometry directly with M13 and M3, with the result that M80's HB is very similar to M13, while M3 lacked the extended blue tail of the other two clusters. Meanwhile, \\citet{GCLSA1999} presented extensive Str\\\"{o}mgren photometry of NGC~6752, M13, and M3, showing extensive blue tails in the former two compared with the latter. Neither M80 nor NGC~6752 had been studied extensively for abundances until \\citet{GratAl2001} and \\citet{GBNF2002} determined [Al/Fe] for 39 stars in NGC~6752 in total, extending the data of \\citet{ND95}. These were significant results because they probed less evolved stars near the main-sequence turnoff, on the subgiant branch, and at the base of the RGB, which are below the point that mixing theories predict that aluminum can be produced. Using a non-LTE analysis, \\citet{GratAl2001} observed dwarfs with [Al/Fe] as low as $-0.76$ dex from the Al~I resonance lines, and subgiants with [Al/Fe] as high as $+0.86$ dex from the doublet at ${\\lambda}{\\lambda}$8773/74~{\\AA}. In fact, the resonance line analysis for the dwarfs yielded [Al/Fe]~$=~-0.18~{\\pm}~0.15$~(s.e.m.)~dex, while the subgiants gave [Al/Fe]~$=~+0.29~{\\pm}~0.11$~dex (s.e.m.). The results for the dwarfs are uncertain due to the difficulty of analyzing resonance lines with non-LTE corrections of as much as $+0.6$~dex; yet, the results are still surprisingly low. What causes the drastic change from the main sequence to the subgiant branch: atmospheric effects, the different choice of lines, or an actual physical phenomenon? The \\citet{GBNF2002} results are more in line with the results of \\citet{ND95} and with other clusters. We discuss implications of the aluminum data in NGC~6752 in more detail in section~\\ref{sec:final_look}. M80, on the other hand, has no published abundances and is in need of further investigation, especially given its similarities to M13 in metallicity and HB morphology. The rest of the paper is outlined according to the following: We begin with a description of the observations in Section~\\ref{sec:obs}, followed by our data reduction techniques in Section~\\ref{sec:ccd}. We then discuss membership criteria in Section~\\ref{sec:rv}. After culling the data, we show the results of our abundance analysis in Section~\\ref{sec:analysis} and give our final conclusions in Section~\\ref{sec:conclude}. ", "conclusions": "\\label{sec:conclude} \\subsection{Summary} \\label{sec:summary} We begin the conclusions by summarizing our results: \\begin{itemize} \\item We observed 21 giants in the globular clusters M80 and NGC~6752 with spectra of sufficient quality to determine abundances. \\item Both M80 and NGC 6752 display a spread of aluminum abundances, having mean abundances of $+0.37$~dex and $+0.51$~dex, respectively. The abundance spreads are 0.43 ({\\stdv}) dex for M80 and 0.36 ({\\stdv}) dex for NGC~6752. \\item The aluminum data cannot resolve the discrepancy in the [Al/Fe] values near the main-sequence turnoff and the subgiants as observed by \\citet{GratAl2001}. \\item No trends between [Al/Fe] ratios and magnitude are discernible in our small sample. \\item The mean [Fe/H] value for M80 is $-1.73$, which is the first spectroscopic determination for this cluster. For NGC~6752, the mean [Fe/H] is $-1.58$, which is consistent with previous results. \\item The Fe-peak elements chromium and nickel follow iron closely. \\item The [Ca/Fe] enhancements are consistent with the ${\\alpha}$-enhancements observed in other clusters. \\item The [Eu/La] ratio is constant for both clusters at $+0.42$ dex, which does not appear unusual; however, both [Eu/Fe] and [La/Fe] are enhanced in M80 relative to NGC~6752. \\end{itemize} \\subsection{A Final Look At Aluminum} \\label{sec:final_look} Now that we've derived aluminum abundances in M80 and NGC~6752, are we any closer to answering whether or not the variations are the result of primordial pollution, mixing processes, or both? Perhaps when combined with the results of \\citet{GBNF2002} we can get some insight into just how complicated this problem really is (we do not include the Gratton data here since they were biased by use of the $c_{1}$ index to select stars.) Given the differences in the quality of the data between the two studies, and the still small number of stars that have been analyzed, it is difficult to draw firm conclusions. If mixing is an ongoing phenomenon, then one would expect that the Al ratio would increase with decreasing magnitude, as it appears to in M13 \\citep{CN2000}; unfortunately, the small numbers make this difficult to discern at this point. An interesting result also comes from both \\citet{GratAl2001} and \\citet{GBNF2002}, who show that stars on the subgiant/lower red giant branch may have just as much Al in them as stars on the upper RGB, where mixing is theoretically possible. These results might suggest that all Al anomalies are the result of pollution, but this raises yet another question: Can it be demonstrated that a globular cluster main sequence star, subgiant, or lower red giant exists that has strongly enhanced Al and is not depleted in C and/or O? That is, if the CNO variations are the result of deep mixing and the Al (and Na) anomalies are primordial, there should be stars that show uncorrelated abundance patterns. But, if the CNO and Na-Al anomalies are both caused by primordial scenarios, then why does evidence exist, in clusters where it's been studied, that the total C$+$N$+$O remains constant from star to star independent of the Na or Al content \\citep{NBS81, CAP88,DCCB91}, and how does one explain $^{12/13}$C ratios near the CN cycle equilibrium value in many other clusters \\citep{SS91,S96b,ZWB96, BSKL97,BSSBN97}? On the other hand, if the anomalies are all created {\\em in-situ}, what is the physics behind them that can occur without completely contradicting well-established basic theories of stellar evolution? It used to be that one could side with one scenario or the other, then the data became more complicated, and one would say that it was somehow a mixture of both primordial and evolutionary scenarios. Maybe in the end that will be the final conclusion, but we are far from proving it; too many questions remain to be answered and too many observations remain to be made." }, "0403/astro-ph0403055_arXiv.txt": { "abstract": "We study the off-equilibrium effects of inflaton on the dynamics of primordial perturbations in the $O(N)$ model. A self-consistent off-equilibrium formalism is employed to investigate the evolution of the inflationary background field and its fluctuations with the back-reaction effects. We find two observable remains left behind the off-equilibrium processes: the running spectral index of primordial density perturbations and the correlations between perturbation modes in phase space, which would serve as the imprints to probe the epoch of inflation, even beyond. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403219_arXiv.txt": { "abstract": "{ We report the first successful application of the astrometric color-induced displacement technique (CID, the displacement of the photocenter between different bandpasses due to a varying contribution of differently colored components to the total light), originally proposed by \\citet{Wielen-1996} for discovering unresolved binary stars. Using the Sloan Digital Sky Survey (SDSS) Data Release 1 with $\\sim 2.5\\,10^6$ stars brighter than 21$^m$ in the $u$ and $g$ bands, we select 419 candidate binary stars with CID greater than 0.5 arcsec. The SDSS colors of the majority of these candidates are consistent with binary systems including a white dwarf and any main sequence star with spectral type later than $\\sim$K7. The astrometric CID method discussed here is complementary to the photometric selection of binary stars in SDSS discussed by \\citet{Smolcic-2004:a}, but there is considerable overlap (15\\%) between the two samples of selected candidates. This overlap testifies both to the physical soundness of both methods, as well as to the astrometric and photometric quality of SDSS data. } ", "introduction": "It is believed that 50\\% of all stars belong to multiple systems \\citep{Heintz-1969:a}. Nevertheless, being aware that a specific star is a binary is always useful because either one throws it out of the sample or updates the model to describe it and begins some follow-up observations! Whether one deals with stellar evolution or galactic dynamics, binaries always receive some special considerations. So, it is important to be able to detect the binary nature of a star at an early stage of an investigation by any possible means. Besides spectroscopy, photometry, and interferometry, astrometry coupled to photometry has lately emerged as a way of revealing the binary nature of a source \\citep{Wielen-1996}. That method relies upon either a change in the position of the source as its brightness varies (Variability-Induced Movers, VIM) or a photometric-band dependence of the position (Color-Induced Displacement, CID). Whereas VIM requires several observations along the brightness variation cycle and only works when at least one component is variable, it only takes one image in each band to identify a CID and can be done for non-variable stars. Despite the number of multi-band photometric surveys, none so far has carried sufficiently accurate astrometry in at least two distinct bands. This lack of observations has been recently alleviated by the Sloan Digital Sky Survey. The Sloan Digital Sky Survey \\citep[SDSS;][ and references therein]{York-2000:a,Abazajian-2003:a} is revolutionizing stellar astronomy by providing homogeneous and deep ($r < 22.5$) photometry in five passbands \\citep[$u$, $g$, $r$, $i$, and $z$;][]{Fukugita-1996:a,Gunn-1998:a,Hogg-2001:a,Smith-2002:b} accurate to 0.02 mag \\citep{Ivezic-2003:a}. Ultimately, up to 10,000 deg$^2$ of sky in the Northern Galactic Cap will be surveyed. The survey sky coverage will result in photometric measurements for over 100 million stars and a similar number of galaxies. Astrometric positions are accurate to better than 0.1 arcsec per coordinate (rms) for point sources with $r<20.5^m$ \\citep{Pier-2003:a}, and the morphological information from the images allows robust star-galaxy separation to $r \\sim$ 21.5$^m$ \\citep{Lupton-2003:a}. Using the SDSS data, we report on the first successful identification of Color-Induced Displacement (hereafter CID) binaries. The underlying ideas of that method are given in Sect.~\\ref{sect:cid}. Sect.~\\ref{sect:simu} describes the simulation that allowed us to optimize the screening of the data described in Sect.~\\ref{sect:data}. In Sect.~\\ref{sect:results}, we present our results and compare them with those of \\citet{Smolcic-2004:a}, who have recently used color selection to identify a stellar locus made of white dwarf+M dwarf binaries. . ", "conclusions": "The color induced displacement method described by \\citet{Wielen-1996} as a way of detecting binaries has been successfully applied to the first public release of the SDSS data. We identify about 400 systems whose changes in position are essentially consistent with a white dwarf coupled to a lower end (later than $\\sim$K7) main-sequence star. We therefore expect $\\sim2\\,000$ CID binaries at the completion of the SDSS observation campaign. This identification of binaries is an independent confirmation of the color based results of \\citet{Smolcic-2004:a}. However, whereas they had a lower bound on $g-r$ of 0.3, the astrometric criterion allows us to identify candidate binaries down to $g-r=-0.4$. On the other hand, color selection they utilized is more sensitive to binaries with angular separations smaller than the sensitivity of the CID method. Though the approach has proven to give results, its efficiency is extremely low. Whereas \\citet{Marchal-2003:a} quote at least 30\\%\\ of binaries among M stars (the percentage grows with the mass of the star along the main-sequence), only 0.02\\%\\ are detected through their CID effect. It is noteworthy that with such a low fraction, the CID binaries do not affect the overall SDSS astrometric precision. Because of their much better astrometric precision (typically a few $\\mu$as), space-based astrometry missions like SIM and Gaia will eventually supersede the SDSS results presented here. According to a Gaia preparatory study \\citep{Arenou-2001:a}, the latter could, for instance, detect a M0 companion to a G0 dwarf star at a $3\\sigma$ level at a separation as low as 2.3 mas. In terms of separations, this is $\\sim 200$ times better than the sensitivity of the CID method applied to the SDSS data." }, "0403/astro-ph0403505_arXiv.txt": { "abstract": "We here report results of an {\\it INTEGRAL} observation of the X--ray burst and atoll source Ser X-1 performed on May 2003. The object was observed for a total of 400 ks but nearly 8$^\\circ$ off-axis due to the amalgamation with an observation of SS 433, the pointing target source. Ser X-1 was detected up to 30 keV with unprecedented positional accuracy for a high-energy emission; a sharp spectral drop is evident beyond this energy. Significant variability is seen in the 20--30 keV light curve. Comparison with previous observations indicates that the source was in its high (banana) state and displayed a soft spectrum during the {\\it INTEGRAL} pointing. A (non simultaneous) broadband radio-to-$\\gamma$--rays broad-band spectral energy distribution for Ser X-1 is also presented for the first time. ", "introduction": "The low-mass X--ray binary Ser X-1, or 4U 1837+04 is known to host a neutron star as the accreting object; it is classified as an Atoll source (e.g., Liu et al. 2001). Archival {\\it EXOSAT} data (Seon \\& Min 2002) showed this object in the banana (i.e. high intensity) state during the observations. More recent {\\it BeppoSAX} and {\\it RXTE} pointing (Oosterbroek et al. 2001) caught the source while it was again in a high activity state, with an unabsorbed flux (1--200 keV) of 8.0$\\times$10$^{-9}$ erg cm$^{-2}$ s$^{-1}$. Its spectrum was well described by a combination of a blackbody-like and Comptonization (Titarchuk 1994) models to account for the observed hard tail. A reflection component could not be excluded, but the data quality could not provide a definitive conclusion. Up to now, however, the X--ray emission observed from Ser X-1 with the most recent high-energy missions was never well representative of the hard (island) state which is generally seen when Atoll sources are undergoing the low intensity phase (Barret 2001). Several X--ray bursts (lasting tens of seconds at most) were also detected from Ser X-1; moreover, during 2001, {\\it BeppoSAX} pinpointed a very long ($\\sim$4 hours) X--ray burst (Cornelisse et al. 2002), making this source join the group of `superbursters' (see Kuulkers 2003 for a review). By studying X--ray bursts observed with {\\it Einstein}, Christian \\& Swank (1997) deduced a distance to the source of 8.4 kpc. This implies a 1--200 keV luminosity of 6.7$\\times$10$^{37}$ erg s$^{-1}$ during the {\\it BeppoSAX} observation of Oosterbroek et al. (2001), which means roughly one third of the Eddington luminosity for a neutron star. The optical counterpart to Ser X-1, located in a crowded stellar field (Thorstensen et al. 1980), was correctly identified by Wachter (1997) and, subsequently, spectroscopically confirmed and studied by Hynes et al. (2004). Very recently, the radio counterpart was discovered with the VLA (Migliari et al. 2004). We here report on a observation of Ser X-1 performed with the INTErnational Gamma--RAy Laboratory ({\\it INTEGRAL}; Winkler et al. 2003) on May 2003, i.e. less than 7 months after the launch of this spacecraft. The high spectral sensitivity of the high-energy instruments onboard this satellite are optimal to study the behaviour of the hard X--ray tail of this source in case of its presence. A more complete analysis of these data can be found in Masetti et al. (2004). ", "conclusions": "" }, "0403/astro-ph0403675_arXiv.txt": { "abstract": "V407 Vul (RX J1914.4+2456) is a candidate double-degenerate binary with a putative 1.756 mHz (9.5 min) orbital frequency. In a previous timing study using archival ROSAT and ASCA data we reported evidence for an increase of this frequency at a rate consistent with expectations for gravitational radiation from a detached ultracompact binary system. Here we report the results of new {\\it Chandra} timing observations which confirm the previous indications of spin-up of the X-ray frequency, and provide much tighter constraints on the frequency derivative, $\\dot\\nu$. We obtained with {\\it Chandra} a total of 90 ksec of exposure in two epochs separated in time by 11.5 months. The total time span of the archival ROSAT, ASCA and new {\\it Chandra} data is now $\\approx 10.5$ years. This more than doubles the interval spanned by the ROSAT and ASCA data alone, providing much greater sensitivity to a frequency derivative. With the addition of the Chandra data an increasing frequency is unavoidable, and the mean $\\dot\\nu$ is $7.0 \\pm 0.8 \\times 10^{-18}$ Hz s$^{-1}$. Although a long-term spin-up trend is confirmed, there is excess variance in the phase timing residuals, perhaps indicative of shorter timescale torque fluctuations or phase instability associated with the source of the X-ray flux. Power spectral searches for periods longward of the 9.5 minute period do not find any significant modulations, however, the sensitivity of searches in this frequency range are somewhat compromised by the dithering of the Chandra attitude. The observed spin-up is of a magnitude consistent with that expected from gravitational radiation decay, however, the factor of $\\approx 3$ variations in flux combined with the timing noise could conceivably result from accretion-induced spin-up of a white dwarf. Continued monitoring to explore correlations of torque with X-ray flux could provide a further test of this hypothesis. ", "introduction": "Ultracompact binary systems could provide a promising means to observe directly the influence of gravitational radiation on orbital evolution. Moreover, such systems would be ideal sources for detection with spaced based gravitational radiation observatories (such as the planned NASA/ESA LISA mission), opening up the possibility for detailed studies of compact interacting binaries. In recent years a pair of candidate ultracompact systems; V407 Vul (also known as RX J1914.4+2456) and RX J0806+1527 (hereafter J0806) have been proposed. These objects were first discovered by ROSAT (Motch et al. 1996; Israel et al. 1999; Beuermann et al. 1999), and initially were suggested to be members of a ``soft'' class of Intermediate Polars (IPs), with the X-ray periods of 569 and 321 s, respectively, representing the putative spin periods of the accreting white dwarfs. Since their discovery extensive follow-up observations have identified the optical counterparts (Ramsay et al. 2000; Israel et al. 2002; Ramsay, Hakala \\& Cropper 2002). Their soft X-ray spectra, the shape of the X-ray modulation, the phasing of the X-ray and optical modulations, the lack of additional longer periods, and the lack of strong optical emission lines have all called into question their IP credentials (for a discussion see Cropper et al. 2003). However, Norton, Haswell \\& Wynn (2004) have argued that an IP interpretation is still plausible if the systems are stream-fed, pole-switching accretors (ie. no accretion disk), and are viewed from a nearly face-on geometry. It was Cropper et al. (1998) who first suggested that V407 Vul might be a double-degenerate compact binary. They proposed a synchronized, magnetic accretor (polar-like) model with accretion powering the X-ray flux. A non-magnetic variant was subsequently proposed by Marsh \\& Steeghs (2002). In this Algol-like model, the accretion stream impacts directly onto a non-magnetic primary, and the spins are not necessarily synchronized with the orbit. An interesting alternative not requiring accretion was proposed by Wu et al. (2002). They suggested a unipolar inductor model, analogous to the Jupiter - Io system (Clarke et al. 1996). If these systems are indeed compact, and thus the observed X-ray period is the orbital period, then an important discriminating factor is the magnitude and sign of the orbital evolution. If the systems are accreting stably from degenerate donors, the expected evolution is for the orbit frequency to decrease. In a previous study we (Strohmayer 2002) used archival ROSAT and ASCA data to explore the evolution of the 1.756 mHz X-ray frequency of V407 Vul over an $\\approx 5$ yr time period, and found evidence for a positive frequency derivative, $\\dot\\nu$, with a magnitude consistent with simple expectations for gravitational radiation induced decay of a circular orbit. Since a measurement of the frequency evolution places severe constraints on possible models, it is crucial to confirm the initial indications of orbital decay and place tighter constraints on $\\dot\\nu$. In this paper we present the results of new {\\it Chandra} observations which confirm an increase in the X-ray frequency and allow us to place much tighter limits on $\\dot\\nu$. In \\S 2 we describe the {\\it Chandra} observations and the data extraction and analysis. In \\S 3 we discuss our phase coherent timing study, and we show that the inclusion of the {\\it Chandra} data conclusively indicates a positive $\\dot\\nu = 7.0 \\pm 0.8 \\times 10^{-18}$ Hz s$^{-1}$. We also discuss the flux variability of the source and excess variance (timing noise) in the phase residuals. In \\S 4 we discuss the implications of our findings for the nature of V407 Vul. We conclude in \\S 5 with a brief summary and goals for future observations. ", "conclusions": "Our new {\\it Chandra} data confirm that the 1.756 mHz X-ray frequency of RX V407 Vul is increasing at a rate of $\\approx 7 \\times 10^{-18}$ Hz s$^{-1}$. Although this rate is consistent with that expected from gravitational radiation losses in a detached ultracompact binary, an IP interpretation in the context of accretion onto a spinning white dwarf cannot yet be strongly ruled out. In some ways the new {\\it Chandra} results have only deepened the mystery surrounding V407 Vul (and by implication, its sister source, RX J0806.3+1527). After a decade of observations we still do not know the nature of the source with any certainty. However, a number of future observations could help provide the solutions. Deep pointings with XMM could probe more sensitively for an unseen orbital period, and further timing observations with {\\it Chandra} will establish constraints on the dependence of torque fluctuations on X-ray flux. Finally, if these do not suffice to crack the mystery, a detection with a spaced-based gravitational radiation observatory (such as the NASA/ESA LISA mission) would provide definitive evidence for an ultracompact system." }, "0403/astro-ph0403396_arXiv.txt": { "abstract": "{Collimated outflows from accreting white dwarfs have a vital role to play in the study of astrophysical jets.} \\addkeyword{Accretion, Accretion Disks} \\addkeyword{Binaries: Symbiotic} \\addkeyword{Stars: Winds, Outflows} \\addkeyword{White Dwarfs} \\begin{document} ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403443_arXiv.txt": { "abstract": "\\xmmn EPIC observations have revealed a bright point-like X-ray source in the nearby Magellanic-type galaxy NGC 55. At the distance of NGC 55, the maximum observed X-ray luminosity of the source, designated as XMMU J001528.9-391319, is $L_x \\sim 1.6 \\times 10^{39}$ erg s$^{-1}$, placing the object in the ultraluminous X-ray source (ULX) regime. The X-ray lightcurve exhibits a variety of features including a significant upward drift over the 60 ks observation. Most notably a series of X-ray dips are apparent with individual dips lasting for typically 100--300 \\,s. Some of these dips reach almost 100 percent diminution of the source flux in the 2.0--4.5 \\,keV band. The EPIC CCD spectra can be modelled with two spectral components, a very soft powerlaw continuum ($\\Gamma \\approx 4$) dominant below 2 keV, plus a multi-colour disc (MCD) component with an inner-disc temperature $kT \\approx 0.8$ \\,keV. The observed upperward drift in the X-ray flux can be attributed to an increase in the level of the MCD component, whilst the normalisation of the powerlaw continuum remains unchanged. The dipping episodes correspond to a loss of signal from both spectral components, although the blocking factor is at least a factor two higher for the MCD component. XMMU J001528.9-391319 can be considered as a candidate black-hole binary (BHB) system. A plausible explanation of the observed temporal and spectral behaviour is that we view the accretion disc close to edge-on and that, during dips, orbiting clumps of obscuring material enter our line of sight and cause significant blocking or scattering of the hard thermal X-rays emitted from the inner disc. In contrast, the more extended source of the soft powerlaw flux is only partially covered by the obscuring matter during the dips. ", "introduction": "\\chandra and \\xmmn provide powerful facilities for studying the X-ray properties of nearby galaxies. A focus of recent work in this area has been the ultraluminous X-ray source (ULX) phenomenon, namely point-like X-ray sources located outside the nucleus of the galaxy with X-ray luminosities apparently in excess of $L_x > 10^{39}$ erg s$^{-1}$ (\\citealt{roberts00}; \\citealt{colbert02}; \\citealt{miller03}). It is entirely plausible that sources with X-ray luminosities at, or just above, this threshold are mass-transfer binaries containing a stellar mass black-hole (3--20 M$_{\\odot}$) radiating at close to the Eddington limit. Supporting evidence is provided by the fact that many ULXs display similar characteristics to those of established black-hole binaries (BHBs) (\\eg \\citealt{makishima00}). However, the nature of the subset of ULXs with X-ray luminosities in excess of a few $\\times 10^{39}$ erg s$^{-1}$ is less certain, since these could be systems harbouring intermediate mass black-holes \\citep{colbert99}, radiating anisotropically \\citep{king01} or possessing truly super-Eddington discs \\citep{begelman02}. It has been estimated that an accreting black-hole, as opposed to a neutron star, is present in at least 10\\% of all bright X-ray binaries (XRBs)\\citep{mcclintock03}. At the present time there are only 18 dynamically-confirmed stellar-mass BHBs, most of which were discovered as X-ray novae in our own galaxy \\citep{mcclintock03}. To this sample we can add a further 20 or more candidate objects which exhibit all the characteristics of black-hole systems \\citep{mcclintock03}. Using \\chandra and {\\it XMM-Newton}, an individual bright binary X-ray source can be studied out to a distance of about 10 Mpc, hence studies of nearby galaxies have the potential for greatly extending our knowledge of luminous XRBs of all types, including black-hole systems. For example, M31 has been a prime target for recent observations (\\eg \\citealt{kong02}; \\citealt{shirey01}; \\citealt{osborne01}), and at least one good BHB candidate has been identified on the basis of its X-ray properties (RX J0042.3+4115; \\citealt{barnard03}). The other major Local Group galaxy, M33, also hosts many discrete X-ray sources (\\eg \\citealt{haberl01}), including the most luminous persistent X-ray source in the Local Group (M33 X-8; \\citealt*{trinchieri88}). This source is another good black-hole candidate, with recent \\chandra observations revealing characteristics consistent with accretion onto a $> 5 M_{\\odot}$ object \\citep{laparola03}. Yet a further example is the discovery of an eclipsing XRB in NGC 253 \\citep{pietsch03}. In the present paper we discuss the properties of the brightest X-ray source detected in NGC 55. This source sits right on the boundary of the ``normal'' binary/ULX categorisation and, on the basis of its X-ray luminosity and spectral properties, is most probably a black-hole system. NGC 55 is a member of the nearby Sculptor group of galaxies, located in the region of the South Galactic Pole at a distance of 1.78 Mpc \\citep{kara03}. It is morphologically similar to the Large Magellanic Cloud but viewed edge-on with its bar pointing almost along the line of sight ($i=90 \\dg$, \\citealt{tully88}). NGC 55 has previously been studied in the X-ray band through \\rosat PSPC (\\citealt*{read97}; \\citealt*{schlegel97}) and HRI \\citep{roberts97} observations. The PSPC data revealed seven bright point-like X-ray sources coincident with the galaxy (\\citealt{schlegel97}). A subsequent re-analysis of the PSPC data, in conjunction with the HRI data, revealed 25 X-ray point sources coincident with, or in close proximity to, the disc of the galaxy \\citep{roberts97}. The \\rosat observations showed one particular object, located $\\sim7'$ to the east of the main bar complex, to be several times brighter than any other X-ray source in the galaxy (Source 7 of \\citealt{schlegel97}; Source 6 of \\citealt{read97}; and Source N55-14 of \\citealt{roberts97}). The \\rosat PSPC data further revealed this source to be spectrally soft (bremsstrahlung temperature $kT \\sim 0.8 - 1.0$ \\,keV or powerlaw photon index $\\Gamma \\sim 3 - 4$) and mildly absorbed (\\nh $\\sim 2 - 4 \\times 10^{21} \\atpcm$) with a derived X-ray luminosity of $\\sim 7 \\times 10^{38} \\ergsec$ in the 0.1 -- 2.4 \\,keV \\rosat band, adopting a distance of 1.78 Mpc. Crucially, both long- and short-term variability were seen, suggesting this is a luminous accretion-powered XRB. Here we revisit this source using new, high quality \\xmmn observations to investigate its spectral and temporal behaviour. \\begin{figure*} \\begin{center} \\scalebox{3}{{\\includegraphics[width=50mm,angle=0]{figure1psxv.ps}}} \\end{center} \\caption{{\\it Left panel:} The \\xmmn image of the NGC 55 field in a broad (0.3--10.0 \\,keV) bandpass. The field centre is at RA $00^{h} 15^{m} 18.0^{s}$, Dec $-39^{\\circ} 13' 33''$ (J2000) and the image size is $40 \\times 40$ arcminute$^2$. The position of XMMU J001528.9-391319 is highlighted by the arrow. {\\it Right panel:} The equivalent optical DSS-2 (red) image with the rotation angle aligned precisely to a north-south projection. The white contour, which is derived from a lightly smoothed version of the X-ray image (using a circular Gaussian mask with $\\sigma = 1$ pixel $ = 4''$), corresponds to a surface brightness of 16 count pixel$^{-1}$. North is up and East is to the left in each case.} \\label{images} \\end{figure*} ", "conclusions": "\\xmmn observations have revealed a luminous XRB in the nearby galaxy NGC 55. On the basis of its X-ray luminosity and X-ray spectral properties we believe this object is most likely a black-hole system. The lightcurve reveals very interesting spectral variability including, most notably, pronounced dips. Future observations may reveal whether these dips show a pattern of occurrence consistent with an underlying orbital period as is the case for the well-studied BHBs GRO J1655-40 and Cyg X-1. However, detailed investigation of the spectral variations which accompany the dips represents a challenge even for the \\xmmn instrumentation." }, "0403/astro-ph0403169_arXiv.txt": { "abstract": "A method is presented for computing the 6-D phase-space density $f(\\Bx,\\Bv)$ and its PDF $v(f)$ in an N-body system. It is based on Delaunay tessellation, yielding $v(f)$ with a fixed smoothing window over a wide $f$ range, independent of the sampling resolution. It is found that in a gravitationally relaxed halo built by hierarchical clustering, $v(f)$ is a robust power law, $v(f) \\propto f^{-2.5 \\pm 0.05}$, over more than 4 decades in $f$, from its virial level to the current resolution limit. This is valid for halos of different sizes in the $\\Lambda$CDM cosmology, indicating insensitivity to the initial-fluctuation power spectrum as long as the small-scale fluctuations were not completely suppressed. By mapping $f$ in position space, we find that the high-$f$ contributions to $v(f)$ come from the ``cold\" subhalos within the parent halo rather than the halo central region and its global spherical profile. The $f$ in subhalos near the halo virial radius is more than 100 times higher than at the halo center, and it decreases gradually with decreasing radius. This indicates phase mixing due to mergers and tidal effects involving puffing up and heating. The phase-space structure provides a sensitive tool for studying the evolution of subhalos during the buildup of halos. One wishes to understand why the substructure adds up to the universal power law in $v(f)$. It seems that the $f^{-2.5}$ behavior is related to the hierarchical clustering process and is not a general result of violent relaxation. ", "introduction": "\\label{sec:intro} Dark-matter halos are the basic entities in which luminous galaxies form and live. They dominate the gravitational potential and have a crucial role in determining the galaxy properties. While many of the systematic features of halo structure and kinematics have been revealed by $N$-body simulations, the origin of these features is still not understood, despite the fact that they are governed by simple Newtonian gravity. The halo density profile $\\rho(r)$ is a typical example. It is found in the simulations to have a robust non-power-law shape (originally Navarro, Frenk \\& White 1997, NFW; Power \\etal 2003; Hayashi \\etal 2004 and references therein), with a log slope of $-3$ at large radii, varying gradually toward $-1$ or even flatter at small radii. The slope shows only a weak sensitivity to the cosmological model and the initial fluctuation power spectrum (e.g. Colin \\etal 2003; Navarro \\etal 2004), indicating that its origin is due to a robust relaxation process rather than specific initial consitions. In particular, violent relaxation (Lynden-Bell 1967) may be involved in shaping up the density profile, but we have no idea why this profile has the specific NFW shape. The properties of the velocity dispersion tensor is another puzzle. The velocity dispersion profile is slightly rising at small radii and slightly falling at large radii but is rather flat overall (Huss, Jain \\& Steinmetz 1999a; 1999b). The profile of the anisotropy parameter $\\beta(r)$ indicates near isotropy at small radii that is developing gradually into more radial orbits at large radii (Colin \\etal 2000). For a sperical system in equilibrium, the $\\sigma(r)$ and $\\beta(r)$ are related to $\\rho(r)$ via the Jeans equation, but it is not at all clear why $\\sigma(r)$ or $\\beta(r)$ have these specific shapes. An interesting attempt to address the origin of the halo profile has been made by Taylor \\& navarro (2001), who measured a poor-man phase-space density profile by $f_{\\rm TN}(r) =\\rho(r)/\\sigma(r)^3$, and found that it displays an approximate power-law behavior, $f_{\\rm TN} \\propto r^{-1.87}$, over more than two decades in $r$. Using the Jeans equation, they showed that this power law permits a whole family of density profiles, and that a limiting case of this family is a profile similar to NFW, but with an asymptotic slope of $-0.75$ as $r \\rightarrow 0$. This scale-free behavior of $f_{\\rm TN}(r)$ is intriguing, and it motivates further studies of halo structure by means of phase-space density. The simulations of the $\\Lambda$CDM cosmology also reveal that the halos are bulit by a rougly self-similar hierarchical clustering process, where smaller building blocks accrete and merge into bigger halos. At every snapshot, every halo contains a substructure of subhalos on top of a smooth halo component that has been tidally stripped from an earlier generation of substructure. Some of the important dynamical processes invloved in this hierarchical halo buildup are understood qualitatively, including dynamical friction, tidal stripping and mergeres. However, a complete understanding of how these processes work in concert to produce the halo structure and kinematics is lacking. Attempts have been made to explain an inner density cusp using toy models of dynamical stripping and tidal effects during the halo buildup by mergers (e.g. Syer \\& White 1998; Dekel, Devor \\& Hetzroni 2003; Dekel \\etal 2003). However, a similar halo density profile seems to be produced also in simulations where substructure has been artificially suppressed (Moore \\etal 1999b; Alvarez, Shapiro \\& Martel 2002), indicating that the process responsible for the origin of this density profile might be a more robust feature of gravity. The issue of halo substructure has become timely both observationally and theoretically. Tidal streams associated with dwarf satellite galaxies are observed in the halos of the Milky Way and M31 and reveal their histories (Ibata \\etal 2001; this proceedings). Gravitational-lenses provide preliminary indicatons for the presence of substructure in halos at the level predicted by the $\\Lambda$CDM scenario (Dalal \\& Kochaneck 2002). In contrast, the observed number density of dwarf galaxies seems to be significantly lower, thus posing a ``missing dwarf problem\" (Klypin \\etal 1999; Moore \\etal 1999a). Also, the ``angular-momentum problem\" of disk galaxies (e.g. Navarro \\& Steinmetz 2000; Bullock \\etal 2001a) is probably associated with the evolution of substructure in halos (Maller \\& Dekel 2002; Maller, Dekel \\& Somerville 2002). While these problems necessarily invlove baryonic proceses, understanding the gravitational evolution of substructure is clearly a prerequisite for solving them. Aiming at the origin of halo structure, we report here on a first attempt by Arad, Dekel \\& Klypin (2004, ADK) to address directly the halo phase-space structure. The fundamental quantity in the dynamical evolution of gravitating systems is the full, 6D, phase-space density $f(\\Bx,\\Bv)$, which intimately relates to the underlying Vlasov equation, and lies behind any (violent) relaxation process that gives rise to the virialized halo structure. Ideally, one would have liked to compute $f$ free of assumptions regarding spherical symmetry, isotropy, or any kind of equilibrium, but computing densities in a 6D space is a non-trivial challenge. The state-of-the-art N-body simulations, with more than million particles per halo, allow for the first time an attempt of this sort. ADK developed a succesful algorithm for measuring $f(\\Bx,\\Bv)$, and studied its relevant properties and the associated systematic and random uncertainties. They then applyied this algorithm to simulated virialized halos in the $\\Lambda$CDM cosmology, and obtained two surprising new results. First, the phase-space volume distribution of $f$ is a universal power-law, valid in all virialized halos that form by hierarchical clustering. Second, this power law is not directly related to the overall density profile, but is rather driven by the halo substructure. We thus learn that $f(\\Bx,\\Bv)$ provides a useful tool for studying the hierarchical buildup of dark-matter halos and the evolution of substructure in them. ", "conclusions": "\\label{sec:conc} It is important to verify that these results are not numerical artifacts. Based on the error analysis and tests with mock datasets, we believe that the $v(f)$ measured by the DTFE algorithm genuinely reflects the true phase-space properties of the given $N$-body system over a broad range of $f$. The question is whether the phase mixing suffered by the subclumps is an artifact of numerical effects such as few-body relaxation, leading to underestimated inner densities and/or overestimated internal velocities (Binney 2003). The apparent agreement between simulations run with different codes and different resolutions is encouraging. In order to specifically address the effect of two-body relaxation, we intend to run twice a simulation of the same halo with the same number of particles but with a different force resolution (ongoing work with F. Stoehr). Assuming that the simulations genuinely reflect the true physical behaviour, the origin of the robust power-law shape of $v(f)$ from the merging substructure becomes a very interesting theoretical issue. As demonstrated in \\S5, a simple model using the mass function and the scaled profiles of the general halo population in the $\\Lambda$CDM scenario does not reproduce the correct power law. This, and the apparent trend of the $f$ spikes with radius, indicate that the structural and kinematical evolution of the subhaloes in the parent halo are important. Studies of tidal heating and stripping may be found useful in this modelling. It would be interesting to follow the phase-space evolution and the contribution to the overall $v(f)$ by a single, highly resolved subhalo, or many of those, as they orbit within the parent halo and approach its center. This may help us understand the nature of the interaction between the parent halo and its subhaloes, and the origin of the $v(f)$ power law (ongoing works with E. Hayashi and with B. Moore). We saw that the power-law behavior of $v(f)$ is limited to the virial regime. It would be interesting to learn how this shape evolves in time as the halo virializes. A preliminary study (to be concluded and reported in another paper) indicates that in the intermediate-$f$ regime the $v(f)$ of a pre-virialized system is significantly flatter than $f^{-2.5}$, while in the high-$f$ regime it drops in a much steeper way. The $f^{-2.5}$ behavior seems to be a feature unique to virialized systems. We learned that in the haloes that are built by hierarchical clustering, the power-law behavior $v(f)\\propto f^{-2.5}$ reflects the halo substructure. It would be interesting to find out whether this power-law behavior actually requires substructure, or it is a more general phenomenon of virialized gravitating systems, valid independently of substructure. One way to answer this question would be to analyse simulated haloes in which all fluctuations of wavelengths smaller than the halo scale were removed, resulting in a smooth halo formed by monolithic collapse, with no apparent substructure in the final configuration. As described in \\S1, such haloes are known to still have NFW-like density profiles in real space, and one wonders whether they also have the magic power-law $v(f)$. There are preliminary indications for a steeper $v(f)$ in this case (Arad, Dekel \\& Moore, in preparation). If confirmed, it would indicate that the $f^{-2.5}$ behavior, while insensitive to the exact slope of the initial power spectrum, is unique to the hierarchical clustering process, and is not a general result of violent relaxation. Our current results are just first hints from what seems to be a promising rich new tool for analysing the dynamics and structure of virialized gravitating systems. The analysis could become even more interesting when applied to haloes including the associated gaseous and stellar components." }, "0403/hep-th0403132_arXiv.txt": { "abstract": "{ The exact computation of asymptotic quasinormal frequencies is a technical problem which involves the analytic continuation of a Schr\\\"odinger--like equation to the complex plane and then performing a method of monodromy matching at the several poles in the plane. While this method was successfully used in asymptotically flat spacetime, as applied to both the Schwarzschild and Reissner--Nordstr\\o m solutions, its extension to non--asymptotically flat spacetimes has not been achieved yet. In this work it is shown how to extend the method to this case, with the explicit analysis of Schwarzschild de Sitter and large Schwarzschild Anti--de Sitter black holes, both in four dimensions. We obtain, for the first time, analytic expressions for the asymptotic quasinormal frequencies of these black hole spacetimes, and our results match previous numerical calculations with great accuracy. We also list some results concerning the general classification of asymptotic quasinormal frequencies in $d$--dimensional spacetimes. } ", "introduction": "A long time has passed since research first focused on analyzing the linear stability of four dimensional black hole solutions in general relativity \\cite{regge-wheeler, zerilli-1}. However, it was not until very recent times that this stability problem was addressed within a $d$--dimensional setting \\cite{kodama-ishibashi-1, kodama-ishibashi-2, kodama-ishibashi-3}. These papers tried to be as exhaustive as possible, studying in detail the perturbation theory of spherically symmetric black holes in $d$--dimensions and allowing for the possibilities of both charge and a background cosmological constant. Having thus acquired a list of stable black hole solutions, the next question to address within this problem are quasinormal modes---the damped oscillations which describe the return to the initial configuration, after the onset of a linear perturbation (see \\cite{nollert, kokkotas-schmidt} for reviews). Besides their natural role in the perturbation theory of general relativity, quasinormal modes have recently been focus of much attention following suggestions that they could have a role to play in the quest for a theory of quantum gravity \\cite{hod, dreyer}. The idea is to look at those special modes which are infinitely damped, and thus do not radiate. It was suggested in \\cite{hod} that an application of Bohr's correspondence principle to these asymptotic quasinormal frequencies could yield new information about quantum gravity, in particular on the quantization of area at a black hole event horizon. It was further suggested in \\cite{dreyer} that asymptotic quasinormal frequencies could help fix certain parameters in loop quantum gravity. Both these suggestions lie deeply on the fact that the real part of the asymptotic quasinormal frequencies is given by the logarithm of an integer number, a fact that was analytically shown to be true, for Schwarzschild black holes in $d$--dimensional spacetime, in \\cite{motl, motl-neitzke}. A question of particular relevance that immediately follows is whether the suggestions in \\cite{hod, dreyer} are universal or are only applicable to the Schwarzschild solution. Given the mentioned analysis of \\cite{kodama-ishibashi-1, kodama-ishibashi-3}, one has at hand all the required information to address this problem and compute asymptotic quasinormal frequencies of $d$--dimensional black holes. A preliminary clue is already present in \\cite{motl-neitzke}, where the analysis of the four dimensional Reissner--Nordstr\\o m solution yielded a negative answer: the asymptotic quasinormal frequencies obeyed a complicated relation which did not seem to have the required form. While extending this result to both the $d$--dimensional and the extremal Reissner--Nordstr\\o m solutions did not pose great obstacles \\cite{natario-schiappa}, an extension of the analytical techniques in \\cite{motl-neitzke} to non--asymptotically flat spacetimes proves to be a greater challenge. It is the goal of this paper to carry out an extension of the techniques in \\cite{motl-neitzke} to non--asymptotically flat spacetimes, with the explicit analysis of Schwarzschild de Sitter and large Schwarzschild Anti--de Sitter black holes, both in four dimensions. The detailed study of these solutions in $d$--dimensions will appear elsewhere \\cite{natario-schiappa}, including charged solutions in asymptotically de Sitter and asymptotically Anti--de Sitter spacetimes, as well as an analysis of the implications of our results on what concerns the proposals of \\cite{hod, dreyer}, dealing with the application of quasinormal modes to quantum gravity. It is important to stress that even if the ideas in \\cite{hod, dreyer} turn out not to be universal, it is still the case that quasinormal frequencies will most likely have a role to play in the quest for a theory of quantum gravity. Indeed, quasinormal frequencies can also be regarded as the poles in the black hole greybody factors which play a pivotal role in the study of Hawking radiation. Furthermore, the monodromy technique introduced in \\cite{motl-neitzke} to analytically compute asymptotic quasinormal frequencies was later extended, in \\cite{neitzke}, so that it can also be used in the computation of asymptotic greybody factors. It was first suggested in \\cite{neitzke} that the results obtained for these asymptotic greybody factors could be of help in identifying the dual conformal field theory which microscopically describes the black hole, and these ideas have been taken one step forward with the recent work of \\cite{krasnov-solodukhin}. It remains to be seen how much asymptotic quasinormal modes and greybody factors can help in understanding quantum gravity. Let us conclude this introduction with some generics concerning quasinormal frequencies (we refer the reader to the upcoming \\cite{natario-schiappa} for a full list of conventions and details). Later, in section 2, we shall compute asymptotic quasinormal frequencies for a Schwarzschild de Sitter black hole in four dimensional spacetime. Our results will also be shown to match earlier numerical computations with great accuracy. In section 3, we shall study large Schwarzschild Anti--de Sitter black holes in four dimensions, and analytically compute their asymptotic quasinormal frequencies. Again, our results match earlier numerical computations to great accuracy. We end with some comments concerning the general classification of asymptotic quasinormal frequencies in $d$--dimensional spacetimes \\cite{natario-schiappa}. For a four dimensional Schwarzschild black hole, one has the asymptotic quasinormal frequencies $$ \\lim_{n \\to + \\infty} \\omega_{n} \\sim [ {\\mathrm{offset}} ] + i n [ {\\mathrm{gap}} ] + {\\mathcal{O}} \\left( \\frac{1}{\\sqrt{n}} \\right), $$ \\noindent where the real part of the offset is the frequency of the emitted radiation, and the gap are the quantized increments in the inverse relaxation time. Here, the gap is given by the surface gravity. One can try to extend this analysis to more general situations and also include spacetimes with two horizons, but then generic results become much harder to obtain \\cite{mmv-1, padmanabhan, mmv-2, choudhury-padmanabhan}. We shall take the time dependence for the perturbation to be $e^{i\\omega t}$, so that ${\\mathbb{I}}{\\mathrm{m}} (\\omega) > 0$ for stable solutions. There is also a reflection symmetry $\\omega \\leftrightarrow - \\bar{\\omega}$ which changes the sign of ${\\mathbb{R}}{\\mathrm{e}} (\\omega)$. In this case, our quasinormal mode conventions are the following (see \\cite{natario-schiappa} for a full list of conventions in $d$--dimensions). The perturbation master equations of \\cite{kodama-ishibashi-1, kodama-ishibashi-3} can be cast in a Schr\\\"odinger--like form as \\begin{equation} \\label{schrodinger} - \\frac{ d^{2} \\Phi_{\\omega}}{dx^{2}} (x) + V (x) \\Phi_{\\omega} (x) = \\omega^{2} \\Phi_{\\omega} (x), \\end{equation} \\noindent where the potential will vary according to the specific case at hand. The boundary conditions are the usual: incoming waves at the black hole horizon and outgoing waves at infinity (or at the cosmological horizon, for the asymptotically de Sitter case)\\footnote{For the asymptotically Anti--de Sitter situation things will be different.}. These can be written as \\begin{eqnarray*} \\Phi_{\\omega} (x) &\\sim& e^{i\\omega x}\\;\\, {\\mathrm{as}}\\;\\, x \\to - \\infty, \\\\ \\Phi_{\\omega} (x) &\\sim& e^{-i\\omega x}\\;\\, {\\mathrm{as}}\\;\\, x \\to + \\infty, \\end{eqnarray*} \\noindent where $x$ is the tortoise coordinate. Indeed, if the metric is chosen as $g = - f(r)\\ dt \\otimes dt + {f(r)}^{-1}\\ dr \\otimes dr + r^{2} d\\Omega_{2}^{2}$, with parameters $M = \\MM$ for the black hole mass and $\\Lambda = 3 \\LL$ for the background cosmological constant, then at any (event or cosmological) horizon, $f(R_{H})=0$. One can expand near the horizon $f(r) \\simeq (r-R_{H}) f'(R_{H}) + \\cdots$, and it follows for the tortoise $$ x \\equiv \\int \\frac{dr}{f(r)} \\simeq \\int \\frac{dr}{(r-R_{H}) f'(R_{H})} = \\frac{1}{f'(R_{H})} \\log (r-R_{H}) \\equiv \\frac{1}{2k_{H}} \\log (r-R_{H}) \\equiv \\frac{1}{4 \\pi T_{H}} \\log (r-R_{H}), $$ \\noindent locally near the chosen horizon. Here $k_{H}$ is the surface gravity and $T_{H}$ is the Hawking temperature. ", "conclusions": "" }, "0403/hep-th0403060_arXiv.txt": { "abstract": "We present a Mathematica package for performing algebraic and numerical computations in cosmological models based on supersymmetric theories. The programs allow for (I) evaluation and study of the properties of a scalar potential in a large class of supergravity models with any number of moduli, an arbitrary superpotential, \\Ka\\ potential, and D-term; (II) numerical solution of a system of scalar and Friedmann equations for the flat FRW universe with any number of scalar moduli and arbitrary moduli space metric. We are using here a simple set of first order differential equations which we derived in a Hamiltonian framework. Using our programs we present some new results: (I) a shift-symmetric potential of the inflationary model with a mobile D3 brane in an internal space with stabilized volume; (II) a KKLT-based dark energy model with the acceleration of the universe due to the evolution of the axion partner of the volume modulus. The gzipped package can be downloaded from \\url{http: //www.stanford.edu/~prok/SuperCosmology/} or from \\url{http: //www.stanford.edu/~rkallosh/SuperCosmology/} ", "introduction": "The studies of the cosmological aspects of supergravity and string theory have a long history, going back to the beginning of the 80's. At present, there is a new wave of interest in the cosmological aspects of string theory. The subject is rather complicated, partly due to the complexity of the analytical study of these theories. For example, to find the F-term part of the effective potential in supergravity, one should specify the expressions for the \\Ka\\ potential $K$ and superpotential $W$. Even with the simplest $K$ and $W$, the computation of potential is tedious, especially if there is more than one superfield, and the resulting expression is hard to analyse. Deriving cosmological consequences from the models based on string theory and supergravity is intricate too. For non-canonical \\Ka\\ potentials (which are the rule rather than the exception) the equations of motion acquire additional velocity-dependent terms whose effects are not easily understood using intuition based on the simplest scalar field models. For the case of one field, one can always reduce the theory to the canonical form, but this method does not work for the description of the simultaneous motion of several different fields, so one should really solve the system of equations keeping the non-canonical kinetic terms throughout. Our Mathematica-based package ``SuperCosmology'' is intended to simplify the study of supergravity potentials and of the cosmological models based on string theory and supergravity. The programs we describe here were used in papers \\cite{Kallosh:2002gf}-\\cite{Hsu:2003cy}, where only the final results of computations were presented. The purpose of this paper is to present the explanation of the programs and methods we used in computations. Also we present some new cosmological models and use them to demonstrate how our package works. The SuperCosmology package consists of two parts. Part~I has the following Mathematica nb-files:\\\\ \\noindent {\\tt SuperPotential.nb}, \\\\ {\\tt SuperPotential\\_KKLT.nb}, \\\\ {\\tt SuperPotential\\_fine\\_tune.nb}, \\\\ {\\tt SuperPotential\\_D3.nb}. \\\\ In {\\tt SuperPotential.nb}, {\\tt SuperPotential\\_KKLT.nb}, {\\tt SuperPotential\\_fine\\_tune.nb} one finds examples from \\cite{Hsu:2003cy}, \\cite{Kachru:2003aw} and \\cite{Kachru:2003sx}, respectively, of using our program ``SuperPotential'' in computation of a scalar potential. In {\\tt SuperPotential\\_D3.nb}, a new example of an inflationary potential with a mobile D3 brane is studied, which is based on the D3/D7 inflationary model investigated before in \\cite{Herdeiro:2001zb}-\\cite{Hsu:2004hi}. \\noindent In part~II, we present the program ``FRW'' used in \\cite{Kallosh:2002gf} for the numerical solution of the Friedmann equations for a system with any number of scalar fields with geometric kinetic terms of the form ${1\\over 2} G_{ij} (\\phi, \\phi^*) \\partial\\phi^i \\partial \\phi^{j}$ specified by a metric $G_{ij} (\\phi)$ on the scalar manifold. Part~II has the following Mathematica nb-files:\\\\ \\noindent {\\tt FRW\\_N2.nb}, \\\\ {\\tt FRW\\_DarkE.nb}, \\\\ {\\tt FRW\\_LateDarkE.nb}. \\\\ In {\\tt FRW\\_N2.nb}, we show a dark energy model based on the N=2 supergravity model \\cite{Fre:2002pd}, \\cite{Kallosh:2002wj} which was also discussed in \\cite{Kallosh:2002gf}. New results on the dark energy model based on the KKLT model \\cite{Kachru:2003aw}, are presented in {\\tt FRW\\_DarkE.nb} and {\\tt FRW\\_LateDarkE.nb}. The examples in {\\tt FRW\\_DarkE.nb} describe the situation when the system has not yet reached the dS minimum (so that the scalars are still moving and $\\Omega_D$ is still increasing). In {\\tt FRW\\_LateDarkE.nb} we study the long term evolution including the time when the dS minimum is reached by the scalars. ", "conclusions": "The SuperCosmology Mathematica package presented in this paper proved to be useful in numerous applications, both in our previous work as well as in the new models described in this paper. The interesting features of the new models are due to the special choice of potentials allowed in supersymmetric theories, and the non-canonical geometric kinetic terms. One of the new models studied here is the KKLT-based model of dark energy. The kinetic term of the model has an $SL(2,R)$-symmetry typical for string theory and supergravity (see \\cite{Horne:1994mi} for earlier studies of cosmology with $SL(2,R)$-symmetry). One can find a change of variables which will bring one of the scalars to the canonical form, $\\rho= \\sigma +i \\alpha$, $\\sigma=e^{\\sqrt{2/3} \\phi}$, however, the other one cannot be canonical: $$L_{ kin}= 3{ \\partial \\rho \\partial \\bar \\rho\\over (\\rho+\\bar \\rho)^2}={1\\over 2} [(\\partial\\phi)^2+ {3\\over 2} e^{-2\\sqrt{2/3} \\phi} (\\partial\\alpha)^2]$$ The KKLT non-perturbative potential depends on the volume modulus $\\sigma$ and on the axion $\\alpha$ and has a complicated profile. We have shown the contour plot of this potential as well as some trajectories of scalar fields in Fig. 2. Consider, for example, the evolution of the model from the point $\\sigma=130$, $\\alpha=15$. This initial point corresponds to the top left (orange) dot on the Fig. \\ref{fig:Fig2}. The solution obtained numerically shows that the volume modulus $\\sigma$ after some initial increase stops and waits, while the axion evolves towards $\\alpha = 0$ and oscillates around it. Then $\\sigma$ moves back and eventually both fields get trapped at the minimum of the potential. The unusual behavior of the fields is explained by an interplay between the potential and non-canonical kinetic terms. From a more general perspective: in the second order equation for the scalars (\\ref{second}) there is an extra term $\\Gamma^i_{jk}\\dot \\phi^j \\dot \\phi^k$ in addition to the standard friction due to the Hubble parameter and also the contribution of the potential depends on the metric, $G^{ij} {\\partial V\\over \\partial \\phi^j}$. That makes it hard to guess, before a numerical solution of equations is found, why some trajectories end up at the minimum of the potential whereas some other trajectories lead to the volume de-compactification. Our package ``SuperCosmology'' is intended to aid in the study of models related to string theory as shown in our examples. We hope it will prove to be useful for further investigations of the interface between string theory, supergravity and cosmology. \\subsection*" }, "0403/astro-ph0403166.txt": { "abstract": "{Granato et al. (2004) have elaborated a physically grounded model exploiting the mutual feedback between star-forming spheroidal galaxies and the active nuclei growing in their cores to overcome, in the framework of the hierarchical clustering scenario for galaxy formation, one of the main challenges facing such scenario, i.e. the fact that massive spheroidal galaxies appear to have formed much earlier and faster than predicted by previous hierarchical models, while the formation process was slower for less massive objects. Adopting the choice by Granato et al. (2004) for the parameters governing the history of the star formation, of chemical abundances and of the gas and dust content of galaxies, we are left with only two, rather constrained, but still adjustable, parameters, affecting the time- and mass-dependent SEDs of spheroidal galaxies. After having complemented the model with a simple phenomenological description of evolutionary properties of starburst, normal late--type galaxies, and of AGNs, we have successfully compared the model with a broad variety of observational data, including deep $K$-band, ISOCAM, ISOPHOT, IRAS, SCUBA, radio counts, and the corresponding redshift distributions, as well as the 1--$1000\\,\\mu$m background spectrum. Special predictions have been made for the especially challenging counts and redshift distributions of EROs. We also present detailed predictions for the GOODS and SWIRE surveys with the Spitzer Space Telescope. We find that the GOODS deep survey at $24\\,\\mu$m and the SWIRE surveys at 70 and $160\\,\\mu$m are likely to be severely confusion limited. The GOODS surveys in the IRAC channels (3.6 to $8\\,\\mu$m), reaching flux limits of a few mJy, are expected to resolve most of the background at these wavelengths, to explore the full passive evolution phase of spheroidal galaxies and most of their active star-forming phase, detecting galaxies up to $z\\simeq 4$ and beyond. A substantial number of high $z$ star-forming spheroidal galaxies should also be detected by the $24\\,\\mu$m SWIRE and GOODS surveys, while the 70 and $160\\,\\mu$m will be particularly useful to study the evolution of such galaxies in the range $1 \\lsim z \\lsim 2$. However, starburst galaxies at $z \\lsim 1$--1.5 are expected to be the dominant population in MIPS channels, except, perhaps, at $160\\,\\mu$m. ", "introduction": "The standard Lambda Cold Dark Matter ($\\Lambda$CDM) cosmology is a well established framework to understand the hierarchical assembly of dark matter (DM) halos. Indeed, it has been remarkably successful in matching the observed large-scale structure. However the complex evolution of the baryonic matter within the potential wells determined by DM halos is still an open issue, both on theoretical and on observational grounds. Full simulations of galaxy formation in a cosmological setting are far beyond present day computational possibilities. Thus, it is necessary to introduce at some level rough parametric prescriptions to deal with the physics of baryons, based on sometimes debatable assumptions (e.g.\\ Binney 2004). A class of such models, known as semi-analytic models, has been extensively compared with the available information on galaxy populations at various redshifts (e.g.\\ Lacey et al.\\ 1993; Kauffmann, White \\& Guiderdoni, 1993; Cole et al.\\ 1994; Kauffmann et al.\\ 1999; Somerville \\& Primack 1999; Cole et al.\\ 2000; Granato et al.\\ 2000; Benson et al.\\ 2003). The general strategy consists in using a subset of observations to calibrate the many model parameters providing a heuristic description of baryonic processes we don't properly understand. Besides encouraging successes, current semi-analytic models have met critical inconsistencies which seems to be deeply linked to the standard recipes and assumptions. These problems are in general related to the properties of elliptical galaxies, such as the color-magnitude and the [$\\alpha$/Fe]-M relations (Cole et al. 2000; Thomas 1999; Thomas et al. 2002), and the statistics of sub-mm and deep IR selected (I- and K-band) samples (Silva 1999; Chapman et al. 2003; Kaviani et al. 2003; Daddi et al. 2004; Kashikawa et al. 2003; Poli et al. 2003; Pozzetti et al. 2003; Somerville et al. 2004; see Cimatti 2003 for a review). However, the general agreement of a broad variety of observational data with the hierarchical scenario and the fact that the observed number of luminous high-redshift galaxies, while substantially higher than predicted by semi-analytic models, is nevertheless consistent with the number of sufficiently massive dark matter halos, indicates that we may not need alternative scenarios, but just some new ingredients. Previous work by our group (Granato et al.\\ 2001; Romano et al.\\ 2002; Granato et al.\\ 2004) suggests that a crucial ingredient is the mutual feedback between spheroidal galaxies and active nuclei at their centers. Granato et al.\\ (2004, henceforth GDS04) presented a detailed physically motivated model for the early co-evolution of the two components, in the framework of the $\\Lambda$CDM cosmology. \\begin{figure}[tbp] \\centering \\includegraphics[width=9truecm]{c15dstd6090.ps} \\includegraphics[width=9truecm]{c15istd6090.ps} \\caption{Differential (upper panel) and integral (lower panel) $15 \\mu$m counts. The solid line is the sum of contributions from spheroids (dot-dashed line; long dashes single out passively evolving spheroids), spirals (short dashes), starburst galaxies (dotted line) and (type 1 + 2) AGN (filled triangles). Data are from Elbaz et al.\\ (1999), Gruppioni et al.\\ (2002).} \\label{c15std} \\end{figure} \\begin{figure}[tbp] \\centering \\includegraphics[width=9truecm]{c175std6090.ps} \\includegraphics[width=9truecm]{nz175std6090_223mjy.ps} \\caption{$170\\,\\mu$m counts (upper panel) and redshift distribution of sources with S$_{170}>223\\,$mJy over an area of $3\\,\\hbox{deg}^{-2}$. The dotted, dashed, dot-dashed lines and the filled triangles show the contributions of starburst, spiral, (star-forming) spheroidal galaxies and AGN, respectively. Data in the upper panel are by Dole et al.\\ (2001). In the lower panel, the thin continuous line with asterisks is the sum of the various contributions and the thick continuous histogram shows the data by Rowan-Robinson et al.\\ (2003).} \\label{c175std} \\end{figure} In this paper, we present a comprehensive comparison of the model with the available data (number counts and redshift distributions) in near-IR (NIR) to sub-mm bands and extensive predictions relevant for surveys such as GOODS and SWIRE, which are being carried out with NASA Spitzer (formerly SIRTF) Observatory. In Sect.~2 we give a short overview of the GDS04 model for spheroidal galaxies, a description of the phenomenological approach adopted to model the evolution of starburst and normal late-type galaxies, and of active galactic nuclei (AGNs). In Sect.~3 we discuss the determination of the two main parameters controlling the time-dependent spectral energy distributions (SEDs) of spheroidal galaxies. In Sect.~4, the model counts and redshift distributions are compared with data from the ISO surveys and follow-up, from the IRAS $60\\,\\mu$m survey, from the SCUBA $850\\,\\mu$m surveys and follow-up, from the radio 1.4 GHz surveys down to sub-mJy flux densities, and from deep K-band surveys and follow-up. In Sect.~5, we present our predictions for Spitzer GOODS and SWIRE surveys. The main conclusions are summarized in Sect.~6. We adopt the following cosmological parameters: $\\Omega_m=0.3$, $\\Omega_\\Lambda=0.7$, $H_0=70$ km s$^{-1}$. \\begin{figure}[tbp] \\centering \\includegraphics[width=9truecm]{nz15std6090_0.1mjydat.ps} \\includegraphics[width=9truecm]{nz15std6090_0.1mjymod.ps} \\caption{Redshift distribution of sources brighter than $0.1\\,$mJy at $15 \\mu$m, within an area of $6\\cdot 10^{-3}\\,\\hbox{deg}^{2}$. In the upper panel the global (spheroids plus spiral and starburst galaxies) redshift distribution predicted by the model (thin continuous line with asterisks) is compared with data by Elbaz et al. (2002, thick solid histogram with error bars) and by Franceschini et al. (2003, three dot-dashed line). In the lower panel we show the various contributions to the global model distribution (again shown by the thin solid histogram with asterisks): starbursts (dots), spirals (dashes), star-forming spheroids (dots-dashes), passively evolving spheroids (long dashes).} \\label{nz15std01} \\end{figure} \\begin{figure}[tbp] \\centering \\includegraphics[width=9truecm]{nz15std6090_1mjydat.ps} \\includegraphics[width=9truecm]{nz15std6090_1mjymod.ps} \\caption{Redshift distribution of sources brighter than $1\\,$mJy at $15 \\mu$m, within an area of $5.46\\,\\hbox{deg}^{-2}$. In the upper panel the global (spheroids plus spirals, starbursts and AGN) model $z$-distribution (thin continuous line with asterisks) is compared with data by Rowan-Robinson et al. (2003, thick continuous line) and Pozzi et al (2003; three dots-dash, scaled to the same area). In the lower panel, the dotted, short-dashed, dot-dashed lines and the filled triangles show the contributions from starburst, spiral, spheroidal galaxies, and AGN respectively, to the global redshift distribution, represented again by the thin continuous line with asterisks. The long-dashed line singles out the contribution of passively evolving spheroids.} \\label{nz15std1} \\end{figure} ", "conclusions": "Granato et al. (2001, 2004) have shown that the mutual feedback between star-forming spheroidal galaxies and the active nuclei growing in their cores can be a key ingredient towards overcoming one of the main challenges facing the hierarchical clustering scenario for galaxy formation, i.e. the fact that the densities of massive high redshift galaxies detected by SCUBA and by deep near-IR surveys are well above the predictions. However, to take full advantage of the wealth of data on extragalactic sources that are rapidly accumulating in the IR to mm region to test evolutionary models and to assess their parameters, we need to deal with complex and poorly understood processes that determine the time-dependent SEDs of the various populations of galaxies. Indeed, semi-analytic models must rely on a large number of adjustable parameters. We have carried out a detailed comparison of the physically grounded GDS04 model, keeping their choice for the parameters controlling the star-formation history, the chemical enrichment and the evolution of dust and gas content of massive spheroidal galaxies. We are therefore left with only two adjustable parameters, affecting their near-IR to mm SED (see Sect.~\\ref{sect:SED}) computed using the code GRASIL that includes a full treatment of star-light reprocessing by dust; as described in Sect.~\\ref{sect:param}, their values are constrained mostly by $15\\,\\mu$m and $K$-band counts. A simplified phenomenological approach (Sect.~\\ref{sec:other}) has been adopted to deal with the other relevant galaxy populations (normal late-type and starburst galaxies), and the contribution by AGN has been estimated by coupling the cosmological evolution of AGN in the X-ray bands with detailed SEDs (Sect.~\\ref{sec:agn}). The model predictions have then been tested against a broad variety of observational data, including, in addition to the $15\\,\\mu$m and $K$-band counts, the redshift distributions of sources brighter than 0.1 and 1 mJy at $15\\,\\mu$m, the SCUBA counts at $850\\,\\mu$m, the available (although still scanty) data on the redshift distribution of sources brighter than 5 mJy at $850\\,\\mu$m, the ISOPHOT 90 and $170\\,\\mu$m counts and the corresponding redshift distributions, the IRAS $60\\,\\mu$m counts, the radio 1.4 GHz counts, the ISOCAM $6.7\\,\\mu$m counts and redshift distribution, the redshift distributions of galaxies to the magnitude limits $K= 20$, 23, and 24, and the 1--$1000\\,\\mu$m background spectrum. Specific predictions for the $K$-band counts and redshift distributions of EROs have been worked out and compared with data. Encouraged by the good agreement of model predictions with all these data sets, we have worked out detailed predictions for the GOODS and SWIRE surveys with the Spitzer Space Telescope. In agreement with previous estimates, we find that the GOODS deep survey at $24\\,\\mu$m and the SWIRE surveys at 70 and $160\\,\\mu$m are likely to be severely confusion limited. The GOODS surveys in the IRAC channels (3.6 to $8\\,\\mu$m), reaching flux limits of a few $\\mu$Jy, are expected to resolve most of the background at these wavelengths, to explore the full passive evolution phase of spheroidal galaxies and most of their active star-forming phase, detecting galaxies up to $z\\simeq 4$ and beyond. A substantial number of high $z$ star-forming spheroidal galaxies should also be detected by the $24\\,\\mu$m SWIRE and GOODS surveys, while the 70 and $160\\,\\mu$m surveys will be particularly useful to study the evolution of such galaxies in the range $1 \\lsim z \\lsim 2$. However, starburst galaxies at $z \\lsim 1$--1.5 are expected to be the dominant population in MIPS channels, except, perhaps, at $160\\,\\mu$m. We plan to apply our model to make predictions for the forthcoming surveys with Herschel, Planck, LMT and ALMA. %%%%%%%%%%%%%%%%%%%%%%%%%%%%" }, "0403/astro-ph0403661_arXiv.txt": { "abstract": "The diffuse galactic gamma-ray spectrum measured by the EGRET experiment \\citep{Hunter:1997} are interpreted within a scenario in which cosmic rays (CRs) are injected by three different kind of sources, (i) supernovae (SN) which explode into the interstellar medium (ISM), (ii) Red Supergiants (RSG), and (iii) Wolf-Rayet stars (WR), where the two latter explode into their pre-SN winds \\citep{Biermann:2001iu,Sina:2001}. ", "introduction": "\\begin{table}[b] \\label{tab_sources} \\centering \\caption{Interaction spectra for the different type of supernovae. $E_{\\rm knee}$ and $E_{\\rm cut-off}$ are in GeV. $\\gamma_1$ and $\\gamma_2$ are the spectral indices ($\\Phi=\\Phi_0 E^\\gamma$) below and above $E_{\\rm knee}$. $Z$ is the charge of the nucleus.} \\vspace{0.3cm} \\begin{tabular}{ccccc} \\hline SN type& $E_{\\rm knee}$ & $E_{\\rm cut-off}$ &$\\gamma_1$& $\\gamma_2$ \\\\ \\hline ISM & & $3\\,Z \\,10^{5}$ & -2.75 & \\\\ RSG & & $3\\,Z \\, 10^{5}$ & -2.33 & \\\\ WR & $2\\,Z \\,10^{6}$& $Z \\, 10^{8}$ & -2.88 & -3.21 \\\\ \\hline \\end{tabular} \\end{table} Conventional models of diffuse galactic gamma-ray production are based on three main processes. Gamma-rays are produced through the decay of $\\pi^0$ as secondary particles of hadronic collisions of CRs with the ISM, such as proton-hydrogen or proton-helium collisions, in bremsstrahlung processes of CR electrons with the ISM, and through inverse Compton scattering of CR electrons with interstellar radiation fields. These models can explain a wide range of observations like the energy spectra below $\\approx 1$~GeV or the integrated flux from the outer parts of our Galaxy. Nonetheless the diffuse galactic gamma-rays observed by the EGRET experiment from the inner Galaxy above $\\approx$~1~GeV exceeds by about 60$\\,$\\% the intensity predicted by these calculations (the measured spectrum is too hard). In this contribution we shall present first results on the diffuse gamma-ray production expected in the model by \\citet{Biermann:2001iu}. In this model, in addition to the aforementioned processes, interactions of hadrons of the relatively hard CR injection spectrum (see Tab.~\\ref{tab_sources}) with the pre-SN winds are considered as further possible sources of diffuse galactic gamma-rays \\citep{Biermann:2001iu}. \\begin{figure}[t] \\centering \\includegraphics[width=7.5cm]{0.05new.ps} \\caption{The log of the mean energy of the CR particles versus the number of shock crossings for $V_{sh}=0.01c$. Starting from the bottom for $5^{o}, 25^{o}, 65^{o}, 80^{o}$ and $85^{o}$ respectively. We see the difference in the energy gain of CR particles for the almost perpendicular case compared to smaller shock inclinations. An effect that \\citep{Jokipii:1987} pointed out as well.} \\label{fig_athina1} \\end{figure} In section~2 a Monte Carlo code is developed to model particle acceleration in non-relativistic near parallel and highly oblique shock configurations, including cross-field diffusion, with application to RSG and WR winds. In section~3 we review the adopted model of diffuse galactic gamma ray production and introduce all necessary input parameters. In section~4 and 5, we present our results, compared to EGRET, CASA-MIA and KASCADE data. In section~6, we summarize our findings and discuss some further improvements in our models. ", "conclusions": "The presented model is very promising in explaining the measured diffuse gamma-ray flux. Also it seems to be capable to reproduce the C/B ratio as well as the antiproton flux \\citep{Sina:2001}. It is quite a plausible scenario in which the CR interact mostly in the environment close to their sources. Some stars, like RSG or WR, provide sufficient material, ejected as powerful winds at the end of their lives, before they explode as supernovae. This wind material provides most of the grammage crossed by CR particles seen at Earth. We expect the data in the TeV range from MAGIC and HESS experiments to allow further insight or put constraints on the present model. More detailed studies are planned to calculate the various predictions of the model by \\citet{Biermann:2001iu} and improve the 3D of the interstellar matter, cosmic ray, and radiation field distributions. \\\\ \\vspace{1cm} \\textit{The present work is being supported by AUGER theory and membership grant 05CU1ERA/3 through DESY/BMBF.} \\vspace{0.5cm}" }, "0403/astro-ph0403511_arXiv.txt": { "abstract": "{ We report on a calculation of the growth of the mass of supermassive black holes at galactic centers from dark matter and Eddington - limited baryonic accretion. Assuming that dark matter halos are made of fermions and harbor compact degenerate Fermi balls of masses from $10^{3}M_{\\odot}$ to $10^{6}M_{\\odot}$, we find that dark matter accretion can boost the mass of seed black holes from about $\\sim 5M_{\\odot}$ to $10^{3-4}M_{\\odot}$ black holes, which then grow by Eddington - limited baryonic accretion to supermassive black holes of $10^{6 \\, - \\, 9}M_{\\odot}$. We then show that the formation of the recently detected supermassive black hole of $3\\times 10^{9}M_{\\odot}$ at a redshift of $z = 6.41$ in the quasar SDSS J114816.64+525150.3 could be understood if the black hole completely consumes the degenerate Fermi ball and then grows by Eddington - limited baryonic accretion. In the context of this model we constrain the dark matter particle masses to be within the range from 12 ${\\rm keV/c}^{2}$ to about 450 ${\\rm keV/c}^{2}$. Finally we investigate the black hole growth dependence on the formation time of the seed BH and on the mass of the seed BH. We find that in order to fit the observed data point of $M_{BH} \\sim 3 \\times 10^{9}M_{\\odot}$ and $z \\sim 6.41$, dark matter accretion cannot start later than about $2 \\times 10^{8}$ years and the seed BH cannot be greater than about $10^{4}M_{\\odot}$. Our results are in full agreement with the WMAP observations that indicate that the first onset of star formation might have occurred at a redshift of $z \\sim 15 \\; - \\; 20$. For other models of dark matter particle masses, corresponding constraints may be derived from the growth of black holes in the center of galaxies. ", "introduction": "Over the past few years, the idea of dark matter (DM) and the possible existence of supermassive black holes (BH) of masses from $10^{6.5}$ to $10^{9.5}M_{\\odot}$ at the center of galaxies (Macchetto et al. \\cite{macchetto97}; Sch\\\"odel et al. \\cite{schodel02}) have become deeply rooted within the astrophysical community. The exploration of the relationship between these two intriguing problems of modern astrophysics has been the subject of an increasing number of papers. It has been established that the mass of the central BH is tightly correlated with the velocity dispersion $\\sigma$ of its host bulge, where it is found that $M_{BH} \\sim \\sigma^{4-5}$ (Faber et al. \\cite{faber97}; Magorrian et al. \\cite{magorian98}; Ferrarese \\& Merritt \\cite{ferrase00}; Gebhardt et al. \\cite{gebhardt00}; Ferrarese \\cite{ferrase02}; Haering \\& Rix \\cite{rix04}). This tight relation between the masses of the BHs and the gravitational potential well that hosts them suggests that the formation and evolution of supermassive BHs and the bulge of the parent galaxy may be closely related, e.g. Wang, Biermann \\& Wandel (\\cite{wang00}). In addition, the recent discovery of high redshift quasars with $z > 6 $ (Fan et al. \\cite{fan01}) implies that the formation of supermassive BHs took place over fewer than $10^{9}$years. In spite of the vast and tantalizing work undertaken on BHs, their genesis and evolution are not well understood (see Rees \\cite{rees84} for a review). Since the discovery of quasars in the early 1960s, it has been suggested that these objects are powered by accretion of gas onto the supermassive BHs of masses $10^{6}-10^{9}M_{\\odot}$ (Lynden-Bell \\cite{bell69}). Two scenarios have been discussed in modeling the growth of BHs. One is that BHs grow out of a low mass `seed' BH through accretion (Rees \\cite{rees84}), and another one is that BHs grow by merging (Barkana et al. \\cite{barkana01}; Wang, Biermann \\& Wandel \\cite{wang00}; Gopal-Krishna, Biermann \\& Wiita \\cite{bk03,bk04}). In a recent paper, Duschl \\& Strittmatter (\\cite{ds04}) have investigated a model for the formation of supermassive black holes using a combination of merging and accretion mechanisms. The purpose of this paper is to study the growth of BHs from dark matter and Eddington-limited baryonic accretion. Also, we would like to obtain some constraints on the DM particle masses. In the past, self gravitating neutrino matter has been suggested as a model for quasars, with neutrino masses in the range $0.2 {\\rm keV} \\stackrel {\\textstyle <} {\\sim} m_{f} \\stackrel {\\textstyle <} {\\sim} 0.5 {\\rm MeV}$ (Markov \\cite{markov64}). Later, neutrino matter was suggested to describe DM in clusters of galaxies and galactic halos with masses in the range of $1 {\\rm eV} \\stackrel {\\textstyle <}{\\sim} m_{f} \\stackrel{\\textstyle <}{\\sim} 25 {\\rm eV}$ (Cowsick \\& McClelland \\cite{cowsik73}; Ruffini \\cite{ruffini80}). More recently, fermion balls (FBs) made of degenerate fermionic matter of $10 \\ {\\rm keV} \\stackrel {\\textstyle <}{\\sim} m_{f} \\stackrel{\\textstyle <} {\\sim} 25 \\ {\\rm keV}$ were suggested as an alternative to supermassive BHs in galaxies (Viollier \\cite{viollier94}; Bili\\'c, Munyaneza \\& Viollier \\cite{bmv99}; Tsiklauri \\& Viollier \\cite{tv98}; Munyaneza \\& Viollier \\cite{mv02}). It has been also suggested that if the Galaxy harbors a supermassive BH, then there should be a density spike in which dark matter (DM) falling towards the center could annihilate and the detection of these annihilation signals could be used as a probe for the nature of DM (Bertone, Silk \\& Sigl \\cite{bertone02}; Gondolo \\& Silk \\cite{gondolo99} and Merritt et al. \\cite{merritt02}). The current belief is that DM particles are bosonic and very massive, i.e. $m_{DM} \\stackrel {\\textstyle >}{\\sim} 1 GeV/c^{2}$. However, the absence of experimental constraints on the weakly interacting massive particles (WIMPs) that probably constitute DM leaves the door open for further investigation of the hidden mass of the Universe. In this paper, we will assume DM to be of fermionic matter and described by a Fermi - Dirac distribution with an energy cutoff in phase space (King \\cite{king66}). We will then explore the limits for the DM particle masses in order to reproduce the mass distribution in the Galaxy (Wilkinson \\& Evans \\cite{evans99}) and then study the growth of a seed BH immersed at the center of the DM distribution in galaxies. We use degenerate FBs at the center of DM halos not as replacements for the BHs but as necessary ingredients to grow the BHs in galactic centers. The resulting distribution of stars around a massive BH was studied in detail in the 1970s and early 1980s in the context of globular clusters (Hills \\cite{hills75}; Frank \\& Rees \\cite{frank76}; Bahcall \\& Wolf \\cite{bahcal76}; Duncan \\& Shapiro \\cite{duncan82}; Shapiro \\cite{shapiro85}). Peebles (\\cite{peeble72}) studied the adiabatic growth of a BH in an isothermal sphere and showed that the BH would alter the matter density to an adiabatic cusp with $\\rho \\sim r^{-3/2}$. A few years later, Young (\\cite{young80}) constructed numerical models that confirmed Peebles' results and showed that the BH induces a tangential anisotropy in the velocity dispersion. In this paper, we use the results of previous calculations on the growth of BHs that are accreting stars (Lightman \\& Shapiro \\cite{lightman78}) to investigate the growth of a BH that accretes DM. Here we assume that the physics driving the formation of the power law cusp in the star - star case is the same as in the case of DM particle orbits being perturbed by molecular clouds. Julian (1967) investigated a similar scenario in which the stellar orbits in our Galaxy were perturbed by molecular clouds to explain the stellar velocity dispersion dependence on the star's age. Moreover, Duncan \\& Wheeler (\\cite{dw80}) investigated the anisotropy of the velocity dispersions of the stars around the BH in M87. Given the density distribution in DM halos, we are interested in establishing how a seed BH would grow by accreting DM. Moreover, the comparison of the growth of the BH from accretion of DM and Eddington - limited baryonic matter would give us another piece of information in the debate surrounding the nature of DM. We therefore investigate how a BH seed of $5 \\ M_{\\odot}$ could grow to a $3 \\times 10^{9}M_{\\odot}$ BH as recently detected in quasar SDSS J1148+5251 at z=6.41 (Willot, McLure \\& Jarvis \\cite{willot03}). Such a seed BH of a typical mass between 5 and 9 $M_{\\odot}$ could have evolved in BH binaries (Podsiadlowski, Rappaport \\& Hau \\cite{pod03}). Here, we note that the seed BH could be an intermediate mass black hole (IMBH) of $10^{3-4}M_{\\odot}$ that might have formed from collisions in dense star-forming regions (Portegies Zwart \\& McMillan \\cite{porte02}, Coleman Miller \\cite{miller03} and van der Marel \\cite{van03} for a review). In fact, Wang \\& Biermann (\\cite{wang98}) have established that a BH would grow exponentially with time accreting baryonic matter as long as the supply lasts. In addition, they were able to reproduce the observed correlation $M_{BH}/M_{sph}$ using standard disk galaxy parameters. Assuming a cusp - like distribution of self-interacting DM (Spergel \\& Steinhardt \\cite{spergel00}), Ostriker (\\cite{ostriker00}) has estimated that BHs could grow to $10^{6}-10^{9}M_{\\odot}$ from DM accretion. For completeness, we note that accretion of DM particles by BHs has recently been studied in Zhao, Haehnelt \\& Rees (\\cite{zhao02}) and Read \\& Gilmore (\\cite{read03}). Cosmological parameters of $H_{0}=70 \\ {\\rm km \\ s^{-1} Mpc^{-1}}$, $\\Omega_{m}=0.3$ and $\\Omega_{\\Lambda}=0.7$ are assumed throughout this paper. In section~2, we establish the main equations to describe DM in galaxies. We then discuss the growth of seed BHs from DM and Eddington - limited baryonic matter accretion in section~3 and conclude with a discussion in section~4. ", "conclusions": "In this paper, we have investigated the growth of a stellar seed BH immersed at the center of DM halos with degenerate FBs of mass from $\\sim 10^{3}M_{\\odot}$ to $\\sim 10^{6}M_{\\odot}$. Using the Pauli exclusion principle, we have established that the BH accretion rate strongly depends on the mass $m_{f}$ of the fermions as $\\dot{M}_{BH}\\sim m_{f}^{4}M_{BH}^{2}$ and thus establish for the first time the relationship between BH growth and fermionic DM. We have shown that in order to fit the DM distribution in the Galaxy with such degenerate cores, the DM particles should be in the $ 12 {\\rm keV/c}^{2} \\stackrel {\\textstyle <}{\\sim} m_{f}\\stackrel {\\textstyle <}{\\sim} 450{\\rm keV/c}^{2}$ mass range. We have shown that such DM masses could be used to fit the distribution of DM in dwarf galaxies. FBs of masses $10^{3 \\; - \\; 6} M_{\\odot}$ could only exist in such galaxies where the density drops off as $1/r^{2}$ at large distances. Dwarf galaxies as well as cluster of galaxies do not host FBs as their data can only be fitted by a non - degenerate Fermi - Dirac distribution of the King type. We have argued that the merging of dwarf galaxies would lead to the formation of galaxies with degenerate FBs. The wideness of the fermion mass in the keV range is due to the FB mass range from $10^{3}$ to $10^{6} M_{\\odot}$ that we have adopted in this paper. Our main assumption is the use of fermions as DM candidates in galaxies. However, if one uses bosons instead of fermions, the range of DM particle masses would of course differ from the one obtained in this paper. The range of fermions used in our paper is in conflict with the Lee-Weinberg lower limit of $\\sim 2 {\\rm GeV/c^{2}}$ on the fermion mass (Lee \\& Weinberg \\cite{weinberg77}). This is due to the fact that the derivation of the Lee-Weinberg limit assumes a freezout from equilibrium distributions. Our simple model uses the chemical potential which allows for non equilibrium distributions and this might modify the Lee-Weinberg argument; this remains to be demonstrated. If we use heavy fermions with masses of about $1 {\\rm GeV/c^{2}}$, then according to equation (\\ref{eq:ov}) the maximum mass allowed for the degenerate Fermi core would only be of about $1M_{\\odot}$, which is not enough to grow a stellar mass BH to $10^{3}M_{\\odot}$ in about $10^{8}$ years. On the other hand, very light fermions i.e. $m_{f} \\stackrel {\\textstyle <}{\\sim} 12 {\\rm keV/c}^{2}$ would generate very massive degenerate FBs of masses greater than $10^{6}M_{\\odot}$ and the BH would grow to $10^{9}M_{\\odot}$ in a very short time, i.e $z << 6.41$. The growth of seed BHs from DM accretion is investigated using the quantum cascade mechanism upon which low angular momentum DM particles at the inner Fermi surface are first consumed by the BH and then due to a high degeneracy pressure, higher angular momentum particles are pushed inwards and the process continues until the entire degenerate FB is consumed by the BH. Moreover, molecular clouds have been used as perturbers of DM particle orbits outside the FB and we have shown that the BH grows faster than the FB. After the BH has consumed the entire FB, it then grows by Eddington - limited baryonic accretion to higher masses of $\\sim 10^{9}M_{\\odot}$ at redshifts $z \\sim 6.41$. We also point out that molecular clouds of mass $10^{10}M_{\\odot}$ have also been detected in the host galaxy of the same quasar at a redshift of $z \\sim 6.41$ (Walter et al. \\cite{walter03}). We have also constrained the possible starting time of accretion, i.e. the time of BH seed formation. From our analysis, the mass of a $3\\times 10^{9}M_{\\odot}$ BH in the quasar SDSS J114816.64+525150.3 at a redshift of $z=6.41$ can be fitted exactly if the accretion process starts at a time of about $2 \\times 10^{8}$ years, which corresponds to the reionization time. The seed BH mass is found to be in the range from a few solar masses up to an upper limit of $\\sim 10^{4}M_{\\odot}$. For a seed BH mass of $ 10^{3\\, - \\, 4}M_{\\odot}$, Eddington baryonic matter accretion would be enough to cause the seed BH to grow into a supermassive BH of $3\\times 10^{9}M_{\\odot}$ mass. The data point at a redshift of $z=6.41$ can be fitted by only Eddington baryonic matter accretion with an efficiency of $\\epsilon \\sim 0.01$. Our model provides a method to find the DM particles mass. If it is found that there is a clear lower mass cut of $10^{3}$ to $10^{6} M_{\\odot}$ in the distribution of BH masses, then this mass can be used for the mass of the FB to find the corresponding mass $m_{f}$ of the fermions which will be in the range of 12 keV to 450 keV. If on the other hand the BH mass distribution is a continuous function, then our model of BH growth with DM will probably be ruled out. The postulated DM particles in this paper were non - relativistic at the decoupling time and are usually called cold dark matter particles (CDM). The latter have to be neutral, stable or quasi-stable and have to weakly interact with ordinary matter. As mentioned in section 2.1 , these particles could be axions which have been investigated by DAMA/NAI (Bernabei et al. \\cite{bernabei01}). The heavier particles of mass above $1 {\\rm GeV/c^{2}}$ could also be of the class of DM candidates named WIMPS (Weakly Interactive Massive particles). However, in the standard model of particle physics, CDM cannot be suitable candidates for particles. Thus, a new window beyond the standard model of particle physics has to accommodate these particles for our model of BH growth to work. The DAMA/NAI experiment (see Bernabei et al. \\cite{bernabei05} for a review) which aims at the verification of the presence of DM particles in the Galactic halo, will be able to confirm whether GeV WIMPS or axions of masses of 12 to 450 keV could exist in nature. Observations have shown that the masses of supermassive BHs at galactic centers correlate with the masses of the host bulges, i.e. $M_{BH} \\approx 0.002 \\ M_{bulge}$ (Haering \\& Rix \\cite{rix04}). This result is obtained in our model as long as the Eddington limited accretion dominates the final growth of the BH. This happens for BH masses of more than about $ 10^{5} M_{\\odot}$ (Wang, Biermann \\& Wandel \\cite{wang00}). Mergers of dwarf galaxies as well as the spinning of BHs would play an important role in the growth of the BHs. The consideration of these two effects will be the subject of further investigations. In addition, it would be of great interest to study the growth of BHs from boson DM particles. Although it has been shown that bosons could provide a good fit to the rotation curves in dwarf galaxies, it is not yet clear whether an analogous mechanism could work in galaxies with a $1/r^{2}$ density fall off. While it takes only $8.4\\times 10^{8}$ years to grow supermassive BHs in most distant quasars, the Galactic center might have grown to its current mass of $\\sim 10^{6}M_{\\odot}$ with only DM accretion in a Hubble time. In the following paper, we will address the growth of the Galactic center BH." }, "0403/astro-ph0403402.txt": { "abstract": "Using a high resolution spectrum of the secondary star in the black hole binary \\mbox{A0620$-$00}, we have derived the stellar parameters and veiling caused by the accretion disk in a consistent way. We have used a $\\chi^{2}$ minimization procedure to explore a grid of 800\\,000 LTE synthetic spectra computed for a plausible range of both stellar and veiling parameters. Adopting the best model parameters found, we have determined atmospheric abundances of Fe, Ca, Ti, Ni and Al. The Fe abundance of the star is $\\mathrm{[Fe/H]}=0.14 \\pm 0.20$. Except for Ca, we found the other elements moderately over-abundant as compared with stars in the solar neighborhood of similar iron content. Taking into account the small orbital separation, the mass transfer rate and the mass of the convection zone of the secondary star, a comparison with element yields in supernova explosion models suggests a possible explosive event with a mass cut comparable to the current mass of the compact object. We have also analyzed the Li abundance, which is unusually high for a star of this spectral type and relatively low mass. ", "introduction": "The system \\mbox{A0620$-$00} (V616 Mon) is a low mass X-ray binary (LMXB) discovered as an eruptive X-ray source by {\\it Ariel V} in August 1975 (Elvis et al., 1975). During the outburst, it brighte\\-ned by 6 magnitudes in the optical and after 15 months it had returned to its quiescent magnitude of $m_{V}=18.3$ mag. Spectroscopic observations during quiescence revealed a K5\\,V--K7\\,V stellar spectrum plus an emission line component from an accretion disk surrounding the compact object (Oke 1977; Murdin et al., 1980). Further optical photometric and spectroscopic studies established the orbital period at $P = 0.323$ d and a secondary radial velocity semiamplitude of $K_2 = 457$ {${\\rm km}\\:{\\rm s}^{-1}$} (McClintock \\& Remillard 1986), which implied a mass function of $f(M) = 3.18 \\pm 0.16$ {$M_\\odot$} and thus firm dynamical evidence for a massive compact object---a black hole---in this system. From measurements of the orbital inclination, the compact object mass was estimated at $\\sim 11$ {$M_\\odot\\; $} and the companion star mass at $\\sim 0.7$ {$M_\\odot\\; $} (Shahbaz et al., 1994; Gelino et al., 2001). Many aspects of the origin and evolution of low mass X-ray binaries (LMXBs) still remain unclear. It is believed that these systems begin as wide binaries with extreme mass ratios and orbital separations of $a \\sim 1000$ {$R_\\odot\\; $} (Portegies Zwart et al., 1997; Kalogera \\& Webbink 1998). After filling its Roche lobe, the massive star engulfs its low mass companion and the latter starts to spiraling in to the massive star's envelope (van den Heuvel \\& Habets 1984; de Kool et al., 1987). A close binary forms if the spiral-in ceases before the low mass companion coalesces with the compact helium core of the primary. The helium core continues its evolution and after SN explosion may turn into a neutron star or a black hole. The system becomes an X-ray binary once the secondary star fills its Roche lobe and begins to transfer matter to the compact object. The spiral-in process could give rise to a naked He core, identified with Wolf--Rayet stars that have lost their envelopes (Woosley et al., 1995). The high mass-loss rate (Chiosi \\& Maeder 1986; Nugis \\& Lamers 2000) of these stars makes difficult to understand the formation of compact objects as massive as the black hole in \\mbox{A0620$-$00} (Meynet \\& Maeder 2003; Woosley et al., 1993). However, if the hydrogen envelope of the massive star is removed at the end of the He core burning phase (the so-called {\\it Case C} mass transfer, Brown et al., 1999), the mass lost by wind in the short-lived ($\\sim 10^4$ yr) supergiant stage will not be large. Convection (Langer 1991) and rotation (Maeder \\& Meynet 2000; Heger et al., 2000) influence the structure and evolution of massive stars and subsequently the uncertainties in the treatment of these parameters limit our understanding of the evolution of the progenitors of compact objects. In addition, uncertainties in various aspects of the supernova explosion models affect the predictions of the final remnant mass and the chemical composition of any ejecta captured by the companion. Among the least known ingredients of these models, we may list: \\begin{itemize} \\item The {\\it mass cut}, i.e. the mass above which the matter is expelled at the time of the supernova explosion and below which it remains locked into the compact remnant. \\item The {\\it amount of fallback} or of the mass which is eventually accreted by the compact core (Woosley \\& Weaver 1995; MacFadyen et al., 2001). \\item Possible {\\it mixing} during the collapse phase (Herant \\& Woosley 1994; Herant et al., 1994; Kifonidis et al., 2000; Fryer \\& Warren 2002). \\item The energy of the supernova explosion (Nakamura et al., 2001). \\item The symmetry of the supernova explosion (MacFadyen \\& Woosley 1999; Maeda et al., 2002). \\end{itemize} With the aim of obtaining information on the link between compact objects and their progenitor stars, Israelian et al. (1999) measured element abundances in the secondary star of the black hole binary Nova Scorpii 1994 (GRO J1655$-$40) and found several $\\alpha$-elements (O, Mg, Si, S, and Ti) enriched by a factor of 6--10. Since these elements cannot be produced in a low mass se\\-con\\-da\\-ry star, this was interpreted as evidence of a supernova event that originated the compact object. Taking into account the supernova yields from explosion models of massive stars, the re\\-la\\-ti\\-ve abundances of these elements suggested that the supernova progenitor was in the mass range 25--40 {$M_\\odot$}. Afterwards, these over-abundances were compared with a variety of supernova models, including standard as well as hypernova models (for various helium star masses, explosion energies, and explosion geometries) and a simple model of the evolution of the binary and the pollution of the secondary (Brown et al., 2000; Podsiadlowski et al., 2002). Additional independent evidence for the existence of a supernova event in this system has also been found by Mirabel et al. (2002). In this paper we analyze the chemical abundances of the secondary star in the LMXB \\mbox{A0620$-$00} with the aim of searching for any evidence of nucleosynthetic products from the progenitor of the compact object. ", "conclusions": "We have obtained a high quality spectrum of the secondary star in \\mbox{A0620$-$00} and derived atmospheric chemical abundances. We have set up a technique that provides a determination of the stellar parameters taking into \\mbox{account} any possible veiling from the accretion disk. We find $T_{\\mathrm{eff}} = 4900 \\pm 150$ K, $\\log g = 4.2 \\pm 0.3$, and a veiling (defined as $F_{\\rm disk}/F_{\\rm cont,star}$) of less than 15 per cent at 5000 {\\AA} and decreasing towards longer wavelengths. Assuming a mass for the secondary of $M_2 = 0.68 \\pm 0.18$ {$M_\\odot$}, the estimated surface gravity leads to a stellar radius of $R_2 = 1.1 \\pm 0.4$ {$R_\\odot\\; $}, consistent with the size of the Roche lobe for the secondary given by Gelino et al. (2001). The abundances of Fe, Ca, Ti, Al, and Ni are slightly higher than solar. The abundance ratios of each element with respect to Fe were compared with these ratios in late-type main sequence metal-rich stars. Moderate anomalies for Ti, Ni, and especially Al have been found. A comparison with element yields from spherically symmetric supernova explosion models suggests that the secondary star captured part of the ejecta from a supernova that also originated the compact object in \\mbox{A0620$-$00}. The abundances can be explained if a progenitor with a $\\sim 14$ {$M_\\odot\\; $} helium core exploded with a mass cut in the range 11--12.5 {$M_\\odot$}, such that no significant amount of iron could escape from the collapse of the inner layers. Elements such as O, Mg, Si, S, and C, with unavailable transitions in our spectrum, will be studied to confirm this scenario. The Li abundance in the secondary star in \\mbox{A0620$-$00} is dramatically enhanced in comparison with field late-type main sequence stars, possibly indicating either that this is a young system ($\\sim0.5$--$2\\times10^8$ yr), or the existence of a Li production-preservation mechanism, such as the $\\alpha$--$\\alpha$ reactions, which have to be tested analyzing the $^7{\\rm Li} / ^6{\\rm Li}$ isotopic ratio using future higher S/N optical spectroscopic observations." }, "0403/astro-ph0403041_arXiv.txt": { "abstract": "We present a measurement of the evolution of the stellar mass function in four redshift bins at $0.4 < z < 1.2$, using a sample of more than 5000 $K$-selected galaxies drawn from the MUNICS (Munich Near-Infrared Cluster Survey) dataset. Our data cover the stellar mass range $10^{10} \\leq M/(\\hMsun) \\leq 10^{12}$. We derive K--band mass--to--light ratios by fitting a grid of composite stellar population models of varying star formation history, age, and dust extinction to BVRIJK photometry. We discuss the evolution of the average mass--to--light ratio as a function of galaxy stellar mass in the K and B bands. We compare our stellar mass function at $z > 0$ to estimates obtained similarly at $z=0$. We find that the mass--to--light ratios in the K--band decline with redshift. This decline is similar for all stellar masses above $10^{10}\\,\\hMsun$. Lower mass galaxies have lower mass--to--light ratios at all redshifts. The stellar mass function evolves significantly to $z = 1.2$. The total normalization decreases by a factor of $\\sim 2$, the characteristic mass (the knee) shifts toward lower masses, and the bright end therefore steepens with redshift. The amount of number density evolution is a strong function of stellar mass, with more massive systems showing faster evolution than less massive systems. We discuss the total stellar mass density of the universe and compare our results to the values from the literature at both lower and higher redshifts. We find that the stellar mass density at $z \\sim 1$ is roughly 50\\% of the local value. Our results imply that the mass assembly of galaxies continues well after $z \\sim 1$. Our data favor a scenario in which the growth of the most massive galaxies is dominated by accretion and merging rather than star formation which plays a larger role in the growth of less massive systems. ", "introduction": "\\label{sec:introduction} The stellar mass content in galaxies as a function of redshift is one of the most fundamental observables in the quest to understand galaxy formation and evolution. It provides information on the coupling between the growth of structure through the collapse and subsequent merging of dark matter halos and the physical processes governing the evolution of the baryonic matter. Stellar mass in galaxies grows by star formation within galactic disks, as well as by accretion and the merging of galaxies. In fact, these two processes are related, because star formation in disks can be triggered or enhanced by tidal interaction in close encounters and by merging events. This interplay between cosmological structure formation and star formation is believed to govern the mass assembly history of galaxies. The stellar mass of a galaxy at a given time is difficult to measure, however. While dynamical mass measures are considered to be most reliable, they measure the total mass of an object. The dark matter and gas contributions (which are a function of galaxy type) have to be removed to obtain the stellar mass. These kinds of measurements depend on model assumptions for the dark matter contribution, are observationally very costly, and have therefore only been possible in the local universe so far. The alternative is to convert the luminosity of a galaxy into a stellar mass by means of a model of its stellar population (derived from photometry or spectroscopy) predicting a mass--to--light ratio (\\ML) in a certain wavelength band. Near--infrared (NIR) luminosities of galaxies are believed to be well suited for this approach, as the \\ML values vary only by a factor of roughly 2 across a wide range of star formation histories (SFHs; see, e.g., \\citealp{RR93,KC98a,BD01}). This compares to a variation of a factor of $\\sim 10$ in the B--band. In addition, the optical regime is strongly affected by dust extinction which becomes negligible in the K band for the vast majority of galaxies \\citep{TPHSVW98}. By correlating photometric properties of disk galaxies with inclination, \\citet{MGH03} found the edge--on to face--on extinction correction to be 0.1~mag in the K band. Measuring the stellar masses of galaxies in the local universe by means of modeling their stellar populations has been re--attempted recently using newly available wide--area galaxy surveys. \\citet{Kauffmannetal03a} used spectroscopic data from the Sloan Digital Sky Survey (SDSS), while \\citet{2dF01} and \\citet{BMKW03} combined NIR photometry from the Two Micron All Sky Survey (2MASS) with optical photometry from the 2dF Galaxy Redshift Survey (2dFGRS) and SDSS, respectively, to derive \\ML values and study the stellar mass function (MF) of galaxies. At $z > 0$, suitable multi--wavelength and redshift data are still sparse. Therefore, the integrated stellar mass density, \\rhosz, has been studied, instead of the stellar MF using the available deep field observations. \\citet{DPFB03} and \\citet{Fontanaetal03} studied \\rhosz\\ in the Hubble Deep Fields (HDFs) over the redshift range $01.5$. Thus, for $\\alpha=0.6$, $\\omega=2.0$; for $\\alpha=+1$, $\\omega=2.6$; and for $\\alpha=+1.5$, $\\omega=4.7$. The $\\alpha=0.6$, $\\omega=2.0$ parameters correspond to the classical values for a constant velocity wind (Wright \\& Barlow 1975; Panagia \\& Felli 1975). Olnon (1975) points out that when the density gradient has a power-law dependence on radius, the slope of the SED will be determined by the value of $\\omega$ at the radius where the optical depth is about unity. That is, the size of the effective radiating surface depends on both the density gradient and frequency. Hartmann \\& Cassinelli (1977) showed that a radial outflow whose velocity is a power-law with radius of index $\\beta$ has a density power-law dependence on radius of $\\beta-2$, resulting in a SED power-law ($S_{\\nu}\\propto\\nu^{2/3}$ for a constant velocity wind). The density power-law index $\\omega$ increases so rapidly with $\\alpha$ that $\\omega$ is improbably large for $\\alpha\\geq +1$ ($\\omega\\geq 2.6$). In light of the observed density structure of HII regions discussed above, it seems unlikely that real HII regions have such steep and well-behaved density structures with radius, especially in the very early stages of evolution expected for HC HII regions. We therefore investigate an alternate possible explanation for the observed radio power-law SEDs of HC HII regions, namely hierarchal clumping of nebular gas. As used here, ``hierarchial clumping'' refers to a region filled with clumps of ionized gas having a range of sizes, temperatures, and optical depths defined by power-law distributions. There need not be a medium in which all the clumps are embedded although such a structure could be accommodated in our analysis. A hierarchial clump distribution is not the same as a fractal distribution which posits clumps within clumps within clumps.\tIn a fractal structure the emergent SED is complicated by the fact that every clump is embedded in clumps of larger size, whereas in a hierarchically clumped structure the main complication arises when clumps begin to shadow other clumps, otherwise one does not have to be concerned with radiation transfer through a myriad of larger clumps. The fact that the interstellar medium (ISM) seems to be clumped on all observed size scales in HII regions, planetary nebulae, and neutral atomic and molecular clouds is a strong motivation for the study presented here. High resolution Hubble Space Telescope (HST) images of the Orion nebula (O'Dell 2001 and references therein) have revealed an array of small-scale, ionized structures down to the resolution limit of the HST. The small scale structures (clumps, filaments, knots, etc) are easiest to recognize in Orion because of its proximity to us, but high resolution observations of other HII regions such as M16 also indicate that they are composed of a complex of many clumps of varying sizes (Hester \\etal\\ 1996). It is unlikely that the clumpy structures in Orion and M16 are unique; rather, they probably indicate that such structure is inherent in all HII regions. Small scale ionized clumps are also seen in planetary nebulae (O'Dell \\etal\\ 2002, 2003). High resolution VLBI observations of Galactic HI absorption toward quasars (Faison \\& Goss 2001; Faison \\etal\\ 1998) show that very small clumps (on the order of a few AU) exist in neutral Galactic HI clouds. Extensive CO observations (e.g., Falgarone \\& Phillips 1996; Elmegreen \\& Falgarone 1996; Falgarone \\etal\\ 1998) have clearly demonstrated the existence of small-scale structures in molecular clouds. The origin of the clumpy structure in the various phases of the ISM is controversial and may have different explanations in different environments. For example, turbulence has been suggested by several authors as the origin of structure in molecular and HI clouds (Elmegreen \\& Falgarone 1996; Falgarone \\etal\\ 1998; Lazarian \\& Pogosyan 2000). However, at least some of the structure in HII regions may be due to hydrodynamical instabilities and/or to pre-existing structure in the natal cloud of an emerging HC HII region. The reason for clumpy structures in the ISM around massive stars is beyond the scope of this paper. Here, the observationally established clumpiness of the ISM plus the presence of extended halos around UC and HC HII regions motivates our analysis of the radio free-free spectra of hierarchically clumpy HC HII regions. In the following section, we derive an analytic expression for the free-free emission from a single spherical clump, and employ the result to consider the radio SEDs from an ensemble of clumps. In \\S 3, the model is applied to the source W49N-B2. A brief discussion of the model and its results appear in \\S 4, and concluding are remarks given in \\S 5. ", "conclusions": "We have used expressions for the radio continuum flux density and optical depth of a single, unresolved, uniform (i.e., temperature and density are constant), spherical clump to calculate the SEDs for an ensemble of many such clumps with a power-law distribution of optical depths. The motivations for this morphology are: (1) the empirical evidence for clumping over a wide range of scale sizes in ionized and neutral atomic and molecular phases of interstellar and circumstellar media; and, (2) the intermediate sloped power-law SEDs observed toward a growing number of HC HII regions. The primary thrust of this investigation was to determine if a power-law SED with slopes intermediate between the optically thick and thin limits of $\\alpha=+2$ and $-0.1$ (where $F_{\\nu}\\propto\\nu^{\\alpha}$) can be understood as a consequence of emission from a hierarchically clumped medium; and, if so, to investigate under what conditions intermediate sloped SEDs are formed and over how large a frequency interval they may occur. We have found that it is possible for an ensemble of clumps with a power-law distribution of optical depths to produce power-law SEDs of intermediate slope over a limited bandwidth. The frequency interval over which an intermediate slope holds is controlled by the range of clump optical depths for an appropriate distribution of clumps N($\\tau$). The greater the range in optical depths, $\\tau_{\\rm max}$ to $\\tau_{\\rm min}$, the broader the bandwidth over which an intermediate slope persists. The slope of the intermediate SED power-law is determined primarily by the parameter $\\gamma$ which specifies the fraction of the nebula that is filled with optically thick clumps at a given frequency. In our models, intermediate power-law slopes only appear for values of $\\gamma$ between 1 and 2, however, we have not explored all possible values of parameter space. We find a good fit to the SED of W49N-B2 using the measured radio parameters for this source from De Pree \\etal\\ (2000) for an ensemble of clumps with $\\gamma = 1.5$. The clump optical depths vary from a maximum of 300 to 0.3 at 7 mm and the mean optical depth at this wavelength is 9.5. The geometric covering factor is $C \\approx 0.15$. The covering factor is small enough to easily produce a low density halo around the dense ionized core of this HC~HII region, for which some observational evidence exists. Power-law SEDs of intermediate slopes result from the additive effect of many individual clumps whose turn-over frequencies occur in sequential order over a limited range in frequency. The primary insight gained from our study of hierarchically clumped nebulae is that the distribution of optical depths of the clump population is solely responsible for determining the continuum shape, with variations in size and temperature of the clumps serving only to modulate the level of the free-free emission. Logical extensions of this work would be to investigate the effects introduced by non-spherical clump morphologies, non-uniform clumps (i.e., variations of temperature and density within clumps), and clump shadowing." }, "0403/astro-ph0403640_arXiv.txt": { "abstract": "We present the model of cosmic rays acceleration at ultrarelativistic subshocks and confront it with the observations of gamma-ray bursts (GRBs) and blazars. We investigate cosmic rays acceleration in shocks with Lorentz factors ($\\gamma$) in the range 3 - 40. We show that fluctuations of the magnetic field downstream of the shock do not play an important role in the acceleration process. Results of numerical simulations for shocks with considered Lorentz factors and perpendicular magnetic field inclinations are presented. We fit the derived particle energy spectral index ($\\sigma$) dependence on fluctuations of the magnetic field upstream and $\\gamma$ with a function. ", "introduction": "The acceleration mechanism which operate at ultrarelativistic shock fronts was discovered by Bednarz \\& Ostrowski (1998). The mechanism is different from the diffusive particle acceleration which is assumed to be suppressed at superluminal shock fronts (Bell 1978; Drury 1983). The effect of the magnetic field direction is important for ultrarelativistic shocks because all of them are superluminal. Medvedev \\& Loeb (1999) have shown that the relativistic two-stream instability naturally generate strong magnetic fields which are parallel to the shock front. Therefore ultrarelativistic shocks have to be superluminal even if one could imagine an external magnetic field with the angle between the upstream field and the shock normal smaller than $\\sim 1/\\gamma$. Ultrarelativistic shocks without any mean magnetic fields cannot be considered as real physical phenomena by the same reason. It have appeared a few papers about a particle acceleration at ultrarelativistic shocks without mean magnetic fields (Kirk et al. 2000; Achterberg et al. 2001; Vietri 2003; Lemoine \\& Pelletier 2003) or with subluminal shock geometry (Ellison \\& Double 2002 - parallel shocks). Their acceleration is similar to the diffusive shock acceleration but it includes anisotropies in the angular distribution upstream of the shock. However, particles in this acceleration are able to return to the shock from downstream to upstream due to large magnetic field fluctuations downstream of the shock or due to subluminal shock geometry as in non-relativistic and mildly relativistic regime. Thus, they have failed to understand the actual ultrarelativistic shock acceleration mechanism because the returning is due to small fluctuations of the magnetic field upstream of the shock (the needed fluctuations decrease when the Lorentz factor of the shock increases) and relativistic effects providing a small change of the particle trajectory in the mean field upstream of the shock to be large as measured downstream. To date, the only numerical calculations performed by Bednarz \\& Ostrowski (1998), Bednarz (2000) and in this paper consider the problem of particle acceleration in ultrarelativistic shocks. It is known for a long time that relativistic shocks occur in regions of efficient acceleration of leptons. The acceleration to non-thermal distributions is observed at hot spots of extragalactic radio sources, in blazars, GRBs and pulsar wind nebulae. In order to account for the presence of these high energy leptons, some authors have tried to find the acceleration mechanism. Begelman \\& Kirk (1990) proposed shock-drift acceleration at relativistic shocks to operate at hot spots of extragalactic radio sources. In the mechanism, particles are accelerated in a single shock crossing by drifting parallel (or anti-parallel) to the electric field. Afterwards, Hoshino et al. (1992) described a process of shock acceleration of positrons to non-thermal distributions devoted to account for the synchrotron radiation observed in the Crab Nebula and hot spots. In the model, the gyrating reflected protons downstream of the shock dissipate their energy in the form of collectively emitted, left-handed magnetosonic waves which are resonantly absorbed by the positrons immediately behind the ion reflection region. The dynamics of the Weibel instability has recently been simulated by several research groups using 3D plasma kinetic code. These simulations confirm both the generation of the magnetic field and the particle acceleration downstream of the shock. The particle energising in Silva et al. (2003) simulations (electron-positron plasma) is due to pitch angle scattering in the produced magnetic field after saturation. The energy stored in the magnetic field is transfered back to the plasma particles. Simulations of Frederiksen et al. (2003) have yielded the energy transfer from protons to leptons similar to Hoshino et al. (1992). In their description, the scattered protons create a fluctuating electric field which tends to equilibrate the energy between protons and electrons. Nishikawa et al. (2003) results suggest that electrons and ions are accelerated in the direction transverse to the shock normal only. All the described mechanisms suffer from small energies the particles are able to acquire. The shock-drift acceleration allow for the energy increase of about ten times. Leptons can receive the energy from protons which is about ten times ($\\sim\\gamma$ times in Frederiksen et al. 2003 does not necessarily depend on $\\gamma$) above the thermal energy downstream of the shock. Silva et al. (2003) simulations have led to the generation of high-energy tails in the distribution function, with energies few times above the thermal energy downstream of the shock. In our model presented below, we apply some of these mechanisms to production of seed particles. ", "conclusions": "" }, "0403/astro-ph0403195_arXiv.txt": { "abstract": "We study the evolution of supernova remnants in the circumstellar medium formed by mass loss from the progenitor star. The properties of this interaction are investigated, and the specific case of a 35 $\\msun$ star is studied in detail. The evolution of the SN shock wave in this case may have a bearing on other SNRs evolving in wind-blown bubbles, especially SN 1987A. ", "introduction": "Type II Supernovae are the remnants of massive stars (M $>$ 8 M$_{\\odot}$). As these stars evolve along the main sequence, they lose a considerable amount of mass, mainly in the form of stellar winds. The properties of this mass loss may vary considerably among different evolutionary stages. The net result of the expelled mass is the formation of circumstellar wind-blown cavities, or bubbles, around the star, bordered by a dense shell. When the star ends its life as a supernova, the resulting shock wave will interact with this circumstellar bubble rather than with the interstellar medium. The evolution of the shock wave, and that of the resulting supernova remnant (SNR), will be different from that in a constant density ambient medium. In this work we study the evolution of supernova remnants in circumstellar wind-blown bubbles. The evolution depends primarily on a single parameter, the ratio of the mass of the shell to that of the ejected material. Various values of this parameter are explored. We then focus on a specific simulation of the medium around a 35 $\\msun$ star, and show how pressure variations within the bubble can cause the shock wave to be corrugated. Different parts of the shock wave collide with the dense shell at different times. Such a situation is reminiscent of the evolution of the shock wave around SN 1987A. ", "conclusions": "" }, "0403/astro-ph0403476_arXiv.txt": { "abstract": "{ Using HST/WFPC2 imaging in F606W (or F450W) and F814W filters, we obtained the color maps in observed frame for 36 distant (0.4$\\,<\\,z\\,<\\,$1.2) luminous infrared galaxies (LIRGs, L$_{\\rm IR}(8-1000\\,\\mu$m) $\\geq$\\,$10^{11}$\\,L$_\\odot$), with average star formation rates of $\\sim$100\\,M$_\\odot$\\,yr$^{-1}$. Stars and compact sources are taken as references to align images after correction of geometric distortion. { This} leads to an alignment accuracy of 0.15\\,pixel, which is a prerequisite for studying the detailed color properties of galaxies with complex morphologies. A new method is developed to quantify the reliability of each pixel in the color map without any bias against very red or blue color regions. Based on analyses of two-dimensional structure and spatially resolved color distribution, we carried out morphological classification for LIRGs. About 36\\% of the LIRGs were classified as disk galaxies and 22\\% as irregulars. Only 6 (17\\%) systems are obvious ongoing major mergers. An upper limit of 58\\% was found for the fraction of mergers in LIRGs with all the possible merging/interacting systems included. Strikingly, the fraction of compact sources is as high as 25\\%, similar to that found in optically selected samples. From their K band luminosities, LIRGs are relatively massive systems, with an average stellar mass of about 1.1$\\times$10$^{11}$\\,M$_\\odot$. They are related to the formation of massive and large disks, from their morphologies and also from the fact that they represent a significant fraction of distant disks selected by their sizes. If sustained at such large rates, their star formation can double their stellar masses in less than 1 Gyr. The compact LIRGs show blue cores, which could be associated with the formation of the central region of these galaxies. We find that all LIRGs are distributed along a sequence which relate their central color to their concentration index. This sequence links compact objects with blue central color to extended ones with relatively red central color, which are closer to the local disks. We suggest that there are many massive disks which are still forming a large fraction of their stellar mass since $z$\\,=\\,1. For most of them, their central parts (bulge?) were formed prior to the formation of their disks. ", "introduction": "The evolution of the cosmic star formation density (CSFD) exhibits the history of the stellar mass assembly averaged over all galaxies. A sharp decline of the CSFD since $z\\,\\sim$\\,1 has been found, whereas large uncertainties still remain at higher redshifts, particularly due to the uncertainties and biases regarding dust extinction (e.g. Madau et al.~\\cite{Madau}; Hammer et al.~\\cite{Hammer97}). Investigations of global stellar mass density as a function of redshift indicate that more than one quarter, probably up to half of the present day stars were formed since $z\\,\\sim$\\,1 (Dickinson et al. \\cite{Dickinson} and references therein). This is in agreement with an integration of the CSFD if the latter accounts for all the light re-radiated at IR wavelengths (Flores et al.~\\cite{Flores}). Hence the star-forming activities since $z\\,\\sim$\\,1 still play an important role in the formation of the galaxy Hubble sequence seen in the local universe. {\\it Hubble Space Telescope} (HST) observations show that the merger rate increases significantly at $z\\,\\sim$\\,1, compared with that in the local universe (Le F$\\grave{\\rm e}$vre et al. \\cite{Fevre}; Conselice et al. \\cite{Conselice}). Such events were claimed to be related to dwarf galaxies while massive systems have almost formed before redshift 1 (Brinchmann \\& Ellis~\\cite{BrinchmannEllis}; Lilly et al. \\cite{Lilly98}; Schade et al. \\cite{Schade}). However, with {\\it Infrared Space Observatory} (ISO) mid-infrared imaging, Flores et al. (\\cite{Flores}) inferred that a substantial fraction of star formation since $z\\,\\sim$\\,1 is associated with the LIRGs. These objects are luminous star-forming galaxies at intermediate redshifts ($z\\,\\sim\\,$0.5 to 1), different from the faint blue galaxy population (Genzel \\& Cesarsky \\cite{Genzel}; Franceschini et al.~\\cite{Franceschini03}). It is widely accepted that merger/interaction is very efficient in pushing gas into nuclear region and triggering violent star formation. Therefore LIRGs are suspected to be merging systems and the evolution of these galaxies is linked to the decline of the merger rate (Elbaz et al. \\cite{Elbaz}). Although HST imaging showed that most of the LIRGs are luminous disk/interacting galaxies (Flores et al.~\\cite{Flores}), systematic investigation of their properties is still required to understand their formation and evolution, as well as link them to the counterparts in the local universe. Morphological classification is essential to revealing the nature of the distant LIRGs. However, at high redshifts, it becomes difficult to classify galaxy morphology securely because, the images of the high-z galaxies suffer from reduced resolution, band-shifting and cosmological surface brightness dimming effects, compared with the local objects. With HST {\\it Wide Field Planet Camera 2} (WFPC2) high resolution imaging in two or more bands, spatially resolved color distribution can be used to investigate the distribution of the stellar population, which is complementary to addressing the appearance in single band. Furthermore, the star-forming regions and dusty regions can stand out in the color map. This is very important to the study of LIRGs, in which these regions are expected to be numerous. Canada-France Redshift Survey (CFRS) fields are among the most studied fields at various wavelengths. Two CFRS fields 0300+00 and 1415+52 had been observed deeply by ISOCAM at 15\\,$\\mu$m (Flores et al. ~\\cite{Flores}; 2004 in preparation) and by HST (Brinchmann et al.~\\cite{Brinchmann}). Aimed at performing detailed analyses of morphology, photometry and color distribution for distant LIRGs, additional HST images through blue and red filters have been taken to complement the color information for the two CFRS fields (PI: Hammer, Prop. 9149). In this work, we present the preliminary results of the color distribution of the distant LIRGs. We correct additional effects in HST images and recenter them accurately, which allow us to access the color maps of complex galaxies. We also implement a method to quantify the signal-to-noise (S/N) ratio of the color image in order to give a reasonable cut for the target area in color maps. This paper is organized as follows. Sect. 2 describes the HST imaging observations and the archive data we adopt. In Sect. 3, we describe the various effects which have to be corrected for aligning images in different WFPC2 filters. In Sect. 4, we describe a method we use to generate the color maps. In Sect. 5, we summarize the morphological properties of the distant LIRGs. The results we obtained of the LIRGs are discussed in Sect. 6. Brief conclusions are given in Sect. 7. Throughout this paper we adopt H$_0$\\,=\\,70\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_{\\rm M}$\\,=\\,0.3 and $\\Omega_\\Lambda$\\,=\\,0.7. Unless specified, we exclude PC chip and the unit of pixel refers to that in WF chips. The bands B$_{450}$, V$_{606}$ and I$_{814}$ refer to HST filters F450W, F606W and F814W, respectively. Vega system is adopted for our photometry. \\begin{table*} \\centering \\caption[]{HST imaging with two bands observations in CFRS fields 0300+00 and 1415+52} \\label{hstlist} \\begin{tabular}{cccccccccc} \\hline \\noalign{\\smallskip} Field & BlueFilter & Total Exp. & N$^{\\mathrm{a}}$ & Dither$^{\\mathrm{b}}$ & RedFilter & Total Exp. & N$^{\\mathrm{a}}$ & Dither$^{\\mathrm{b}}$ & Prop.ID$^{\\mathrm{c}}$\\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} 030226+001348 & F606W & 6400 & 5 & 20& F814W & 6000 & 5 & 20& 9149 \\\\ 030227+000704 & F450W & 7000 & 5 & 20& F814W & 6700 & 5 & 20& 6556,5996 \\\\ 030233+001255 & F450W & 6600 & 6 & 0 & F814W & 6400 & 6 & 0 & 5449 \\\\ 030237+001414 & F606W & 6400 & 5 & 20& F814W & 6400 & 5 & 20& 9149 \\\\ 030240+000940 & F606W & 6400 & 5 & 20& F814W & 7000 & 5 & 12.5& 9149,8162 \\\\ 030243+001324 & F450W & 6600 & 6 & 0 & F814W & 6400 & 6 & 0 & 5449 \\\\ 030250+001000 & F606W & 6400 & 5 & 20& F814W & 7000 & 5 & 12.5& 9149,8162 \\\\ 141743+523025 & F450W & 7800 & 6 & 0 & F814W & 7400 & 6 & 0 & 5449 \\\\ 141803+522755 & F606W & 6400 & 5 & 20& F814W & 6800 & 5 & 20& 9149 \\\\ 141809+523015 & F450W & 7800 & 6 & 0 & F814W & 7400 & 6 & 0 & 5449 \\\\ 141724+522512 & F606W & 2800 & 4 & 0 & F814W & 4400 & 4 & 0 & 5090 \\\\ 141731+522622 & F606W & 2800 & 4 & 0 & F814W & 4400 & 4 & 0 & 5090 \\\\ 141737+522731 & F606W & 2800 & 4 & 0 & F814W & 4400 & 4 & 0 & 5090 \\\\ 141750+522951 & F606W & 2800 & 4 & 0 & F814W & 4400 & 4 & 0 & 5090 \\\\ 141743+522841 & F606W &24400 &12 & 0 & F814W &25200 &12 & 0 & 5109 \\\\ 141757+523101 & F606W & 2800 & 4 & 0 & F814W & 4400 & 4 & 0 & 5090 \\\\ 141803+523211 & F606W & 2800 & 4 & 0 & F814W & 4400 & 4 & 0 & 5090 \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\begin{list}{}{} \\item[$^{\\mathrm{a}}$] Number of exposures. \\item[$^{\\mathrm{b}}$] Dither offset among the consecutive exposures, aimed to remove cosmic-rays and hot/bad pixels. Here the largest offset in unit of pixel along x axis is present. \\item[$^{\\mathrm{c}}$] HST Proposal ID. One ID means that both the blue and the red observations were carried out in the same proposal. Two IDs refer to the proposals that the blue and the red observations were observed respectively. \\end{list} \\end{table*} ", "conclusions": "Specific efforts were made to obtain the color maps for galaxies with complex morphology, including the accurate alignment of the blue and red band images and a new method to quantitatively determine the reliability of each pixel in the color map. These efforts allow us to access the spatially resolved color distribution of the distant LIRGs, which often have complex morphologies, relating to interactions/mergers. In two 10$\\arcmin$$\\times$10$\\arcmin$ CFRS fields 0300+00 and 1415+52, HST WFPC2 imaging in F606W (or F450W) and F814W filters is available for 87 square arcminute area. From these fields , we select a representative sample of 36 distant ($0.4\\,<\\,z\\,<\\,1.2$) LIRGs detected in deep ISOCAM observations. Two-dimensional structure analysis is carried out using GIM2D software. With structure parameters and color distribution, a careful morphological classification was performed for the distant LIRGs. We find that about 36\\% LIRGs are spiral galaxies and about 25\\% LIRGs show compact morphology. About 22\\% LIRGs are classified as irregular galaxies, showing complex and clumpy structures. Among 36 LIRGs, only 6 (17\\%) of them were undergoing a major merger episode, revealed by distinctive close galaxy pair with distorted morphology and apparent tidal tails. The fraction of mergers could reach 58\\% if all of the possible post-mergers/pre-mergers are included. Inspection of their stellar masses derived from K band absolute magnitude evidences that LIRGs are massive systems. The LIRGs classified as disk galaxies indeed belong to the large disk galaxy population, and become a significant fraction of large distant disks selected by their sizes. We find that LIRGs are distributed along a sequence in the central color versus compactness diagram. The sequence links the compact LIRGs with relatively blue central color to that of extended LIRGs with central color and compactness close to those of the local normal galaxies. The compact LIRGs showing blue central color are suggested to be the systems forming their bulges, in agreement with the suggestion of Hammer et al. (\\cite{Hammer01}). We argue that the sequence suggests that distant compact LIRGs would eventually evolve into the spiral galaxies in the local universe." }, "0403/astro-ph0403706_arXiv.txt": { "abstract": "The \\al radionuclide can be detected through its decay emission line at 1.809 MeV, as was first observed by Mahoney et al. (1982). Since then, COMPTEL on board of the CGRO satellite, performed a sky survey in this energy range, and provided maps of the \\al distribution in the Galaxy. These results revealed that the main contributors to the synthesis of \\al are most likely the massive stars, which contribute through their winds (Wolf-Rayet stars) and through their supernova explosion.\\\\ Comparison between these observations (in particular observations in localized regions such as the Vela region and the Cygnus region) and the models available at that moment, showed however the need for improvements from both theoretical and observational points of view, in order to improve our understanding of the \\al galactic distribution as well as that of its synthesis.\\\\ With the launch of the INTEGRAL satellite in October 2002, the observational part will hopefully be improved, and the construction of better resolution maps at 1.809 MeV is one of the main aims of the mission. From a theoretical point of view, we need the most up-to-date predictions in order to be able to interpret the forthcoming data.\\\\ In this paper, we address this latter part, and present new results for \\al production by rotating Wolf-Rayet stars and their contribution to the total amount observed in the Galaxy. ", "introduction": "\\al is a radioactive nuclide that can be produced by hydrostatic nucleosynthesis in H burning regions, by explosive nucleosynthesis and by spallation. Its half--life time in its ground state is of $7.2 \\times 10^5$ yr, and it is thus a good tracer of recent nucleosynthetic events in the Galaxy.\\\\ The maps obtained with COMPTEL (Diehl et al. 1995, Kn\\\"odlseder et al. 1999, Pl\\\"uschke et al. 2001) allowed to pin down the main contributors to the diffuse emission observed mainly in the galactic plane : massive stars (Prantzos \\& Diehl 1996, Kn\\\"odlseder 1999). These objects end their lives with a supernova explosion, during which \\al can be synthesized and will eventually be expelled (Heger et al. 2003). The more massive (and shorter lived) ones will also contribute during quiescent evolutionary phases through their winds. ", "conclusions": "We have presented new results concerning \\al production by WR stars from a new grid of models including rotation and updated physics, in particular the most recent prescriptions for mass loss rates.\\\\ Taking rotation into account globally leads to an enhancement of wind ejected mass of \\al by very massive stars.\\\\ Convolved with appropriate IMF and star formation rate indicator, these yields lead to a total mass of about 1.3 \\msun~of \\al originating from WR stellar winds. This value is in agreement with the conclusions drawn by Prantzos (2004) from recent measurements of the line flux ratio $^{60}{\\rm Fe}/^{26}{\\rm Al}$." }, "0403/astro-ph0403256_arXiv.txt": { "abstract": "We use the statistics of strong gravitational lensing based on the Cosmic Lens All-Sky Survey (CLASS) data to constrain cosmological parameters in a spatially-flat, inverse power-law potential energy density, scalar-field dark energy cosmological model. The lensing-based constraints are consistent with, but weaker than, those derived from Type~Ia supernova redshift-magnitude data, and mildly favor the Einstein cosmological constant limit of this dark energy model. ", "introduction": "Recent cosmological measurements strengthen the evidence from Type~Ia supernova redshift-magnitude measurements (Riess et al.~1998; Perlmutter et al.~1999) that the energy density of the current universe is dominated by Einstein's cosmological constant $\\Lambda$, or by a dark energy term in the cosmic stress-energy tensor that only varies slowly with time and space and so acts like $\\Lambda$. These measurements include: (1) more recent Type~Ia supernova redshift-magnitude measurements (see, e.g., Knop et al.~2003; Barris et al.~2004); (2) the space-based Wilkinson Microwave Anisotropy Probe (WMAP) measurement of the cosmic microwave background (CMB) anisotropy, with some input from other measurements (see, e.g., Page et al.~2003; Spergel et al.~2003; Scranton et al.\\ 2003); and (3) other measurements of CMB anisotropy, which indicate the universe is close to spatially flat (see, e.g., Podariu et al.~2001b; Durrer, Novosyadlyj, \\& Apunevych 2003; Melchiorri \\& \\\"Odman 2003), in combination with the continuing strong evidence for low non-relativistic matter density (Chen \\& Ratra 2003b and references therein). See Peebles \\& Ratra (2003), Padmanabhan (2003), Bernardeau (2003), Steinhardt (2003), and Carroll (2004) for reviews of the current state of affairs.\\footnote{ Specific dark energy models and observational measurements are considered in Munshi, Porciani, \\& Wang (2003), Barreiro et al.~(2003), Mainini et al.~(2003), Lima, Cunha, \\& Alcaniz (2003), Silva \\& Bertolami (2003), Amendola et al.~(2003), Linder \\& Jenkins (2003), Makler, Oliveira, \\& Waga (2003), Bean \\& Dor\\'e (2003), {\\L}okas, Bode, \\& Hoffman (2003), Alam et al.~(2003), Choudhury \\& Padmanabhan (2003), Zhu \\& Fujimoto (2004), and Macci\\'o (2004), from which the earlier literature may be accessed.} While Einstein's $\\Lambda$ was the first example of dark energy, nowadays much attention is focused on scalar field models in which the energy density slowly decreases with time and so behaves like a time-variable $\\Lambda$ (see, e.g., Peebles 1984; Peebles \\& Ratra 1988, 2003; Padmanabhan 2003; Steinhardt 2003; Carroll 2004). A simple scalar field dark energy model has scalar field $(\\phi)$ potential energy density $V(\\phi) \\propto \\phi^{-\\alpha}$ at low redshift, with $\\alpha > 0$ (see, e.g., Peebles \\& Ratra 1988; Ratra \\& Peebles 1988). Podariu \\& Ratra (2000), Waga \\& Frieman (2000), and Gott et al.~(2001) examine constraints on this model using Type~Ia supernova redshift-magnitude data. They find that a broad range of $\\alpha$ is consistent with the supernova data.\\footnote{ The proposed SNAP space mission (see http://snap.lbl.gov/, and Schubnell 2003 and Annis et al.~2003) will provide significantly tighter constraints on such models (Podariu, Nugent, \\& Ratra 2001a; Ericksson \\& Amanullah 2002; Caresia, Matarrese, \\& Moscardini 2003; Wang \\& Mukherjee 2003, and references therein). Mukherjee et al.~(2003a, 2003b), Spergel et al.~(2003), Caldwell \\& Doran (2003), Weller \\& Lewis (2003), Giovi, Baccigalupi, \\& Perrotta (2003), and references therein, discuss constraints on scalar field and related dark energy models from CMB anisotropy measurements; upcoming WMAP and other CMB data will improve these constraints.} It is important that these dark energy models be tested by other independent methods. The redshift--angular-size test is one option. Indications from current data, while not as compelling as those discussed above, are consistent with a significant dark energy density at low redshift (see, e.g., Daly \\& Guerra 2002; Zhu \\& Fujimoto 2002; Chen \\& Ratra 2003a; Podariu et al.~2003; Jain, Dev, \\& Alcaniz 2003; Jackson 2003). Future higher-quality data should turn this into a much more precise cosmological test. The redshift-counts test also appears to be on the verge of becoming a very promising test (see, e.g., Newman \\& Davis 2000; Huterer \\& Turner 2001; Podariu \\& Ratra 2001; Levine, Schulz, \\& White 2002). Statistical analyses of strong gravitational lensing can be used to provide constraints on cosmological parameters. Fukugita, Futamase, \\& Kasai (1990) and Turner (1990) note that the rate of gravitational lensing increases rapidly with increasing $\\Lambda$. Ratra \\& Quillen (1992) and Waga \\& Frieman (2000) study gravitational lensing in the inverse power-law potential scalar field dark energy model. The recently completed Cosmic Lens All-Sky Survey (CLASS) is the largest uniform survey for strong lensing (Myers et al.\\ 2003; Browne et al.\\ 2003). The survey has discovered 22 cases of multiple-imaging (that are induced by galaxy-scale lens potentials) out of $\\sim 16,500$ extragalactic radio sources. A subsample of 8958 sources containing 13 multiply-imaged sources satisfy well-defined observational selection criteria and is referred to as the CLASS statistical sample (Browne et al.\\ 2003). The CLASS statistical sample has been used to constrain cosmological parameters (see, e.g., Chae et al.\\ 2002; Chae 2003; Kuhlen, Keeton, \\& Madau 2004; Mitchell et al.\\ 2004) as well as to constrain global properties of galaxies (Chae 2003; Davis, Huterer, \\& Krauss 2003) and galaxy evolution (Chae \\& Mao 2003). The lensing-based constraints on cosmological parameters are consistent with those based on Type~Ia supernova magnitude-redshift data but have larger statistical errors. In this work we use the CLASS statistical sample to constrain the inverse power-law potential scalar-field dark energy model (Peebles \\& Ratra 1988). In linear perturbation theory, a scalar field is mathematically equivalent to a fluid with time-dependent equation of state parameter $w = p/\\rho$ and speed of sound squared $c_s^2 = \\dot p/\\dot\\rho$, where $p$ and $\\rho$ are the pressure and energy density, and the dot denotes a time derivative (see, e.g., Ratra 1991). The XCDM parametrization of this dark energy model approximates $w$ as a constant, which is accurate during the radiation and matter dominated epochs but not in the current, scalar-field dark energy dominated epoch. This XCDM approximation thus leads to inaccurate predictions for the gravitational lensing considered here, which probes the low redshift universe. We emphasize, however, that unlike a lot of earlier work, we do not work in the XCDM approximation, instead we explicitly integrate the scalar-field dark energy equations of motion. In $\\S$2 we summarize the data and method used. Results are presented and discussed in $\\S$3. ", "conclusions": "Figure 1 shows the CLASS lensing-based constraints on the parameters of the spatially-flat inverse power-law potential scalar field dark energy cosmological model. The likelihood is maximized for $\\Omega_{\\rm m,0} = 0.34$ and $\\alpha = 0$, i.e., a conventional cosmological constant. At 68\\% confidence, $\\alpha < 2.7$ and $0.18 < \\Omega_{\\rm m,0} < 0.62$.\\footnote{ If the SDSS measured VF (Sheth et al.\\ 2003; Mitchell et al.\\ 2004) were used instead of the SSRS2 inferred VF (Chae 2003), these ranges would be narrower and the maximum likelihood estimate of $\\Omega_{\\rm m,0}$ would be $\\sim 0.2$. See \\S 2 and Chae (2004, in preparation) for why we choose to use the SSRS2 VF.} However, at 95\\% confidence both $\\alpha = 8$ and $\\Omega_{\\rm m,0}=1$ are allowed. As mentioned in \\S 2, these results are based on the assumption that the comoving number density of early-type galaxies is unchanged from $z \\sim 1$. However, if there were fewer early-type galaxies at intermediate redshifts compared with the present epoch, the maximum likelihood estimate of $\\Omega_{\\rm m,0}$ would become lower and the confidence ranges for $\\alpha$ and $\\Omega_{\\rm m,0}$ would become narrower. The above results are consistent with, but not as constraining as, those derived from Type~Ia supernova redshift-magnitude data (Podariu \\& Ratra 2000; Waga \\& Frieman 2000). They are also consistent with, but more constraining than, those determined using measurements of angular size as a function of redshift (Chen \\& Ratra 2003a; Podariu et al.~2003). It is interesting to note that various disparate data sets give consistent constraints on the inverse power-law potential energy density scalar-field dark energy model that weakly favor the conventional cosmological constant over a dynamical scalar field dark energy. However, current results are tentative and future much larger data sets are required to resolve this issue. Future lensing data (e.g.,\\ CLASS2; see \\S 6 of Chae 2003) would be valuable in this respect and are eagerly anticipated. \\bigskip We thank the anonymous referee for constructive comments. KHC and DWL acknowledge support from the ARCSEC of KOSEF. GC and BR acknowledge support from NSF CAREER grant AST-9875031 and DOE EPSCoR grant DE-FG02-00ER45824." }, "0403/astro-ph0403583_arXiv.txt": { "abstract": "Primordial vector modes describe vortical fluid perturbations in the early universe. A regular solution exists with constant non-zero radiation vorticities on super-horizon scales. Baryons are tightly coupled to the photons, and the baryon velocity only decays by an order unity factor by recombination, leading to an observable CMB anisotropy signature via the Doppler effect. There is also a large B-mode CMB polarization signal, with significant power on scales larger than $l\\sim 2000$. This B-mode signature is distinct from that expected from tensor modes or gravitational lensing, and makes a primordial vector to scalar mode power ratio $\\sim 10^{-6}$ detectable. Future observations aimed at detecting large scale $B$-modes from gravitational waves will also be sensitive to regular vector modes at around this level. ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403060_arXiv.txt": { "abstract": "The second USNO CCD Astrograph Catalog, UCAC2 was released in July 2003. Positions and proper motions for 48,330,571 sources (mostly stars) are available on 3 CDs, supplemented with 2MASS photometry for 99.5\\% of the sources. The catalog covers the sky area from $-90^{\\circ}$ to $+40^{\\circ}$ degrees declination, going up to $+52^{\\circ}$ in some areas; this completely supersedes the UCAC1 released in 2001. Current epoch positions are obtained from observations with the USNO 8-inch Twin Astrograph equipped with a 4k CCD camera. The precision of the positions are 15 to 70 mas, depending on magnitude, with estimated systematic errors of 10 mas or below. Proper motions are derived by utilizing over 140 ground-and space-based catalogs, including Hipparcos/Tycho, the AC2000.2, as well as yet unpublished re-measures of the AGK2 plates and scans from the NPM and SPM plates. Proper motion errors are about 1 to 3 mas/yr for stars to 12th magnitude, and about 4 to 7 mas/yr for fainter stars to 16th magnitude. The observational data, astrometric reductions, results, and important information for the users of this catalog are presented. ", "introduction": "The U.S.~Naval Observatory (USNO) operates the 8-inch Twin Astrograph (\\citet{DH}; \\citet{TACH}) currently from its Flagstaff station (NOFS). The program currently underway with this instrument is the USNO CCD Astrograph Catalog project; it is an ongoing program which started at the Cerro Tololo Interamerican Observatory (CTIO) in 1998. Completion of the all-sky, astrometric observations are expected in May 2004. This second data release, the UCAC2, is a substantial increase in data volume and includes improvements of reduction techniques over the first release, UCAC1, as described in Paper I \\citep{U1}. The UCAC2 encompasses the entire area of UCAC1 and completely supersedes it. The sky coverage has about doubled and new measurements were obtained from early epoch plates that are used in UCAC2 to significantly improve the proper motions with respect to the UCAC1 release. The goal of this project is the densification of the reference frame at optical wavelengths; see also \\citet{spie}. Toward this goal, UCAC provides about a factor of 30 more stars per square degree than the Tycho-2 catalog. The precision of the UCAC observed positions comes close to the precision of Hipparcos positions at current epochs, and surpasses the precision of Tycho-2 positions at about 10th magnitude and fainter. The UCAC2, a compiled catalog, includes Hipparcos and Tycho observational data as well as virtually all ground-based catalogs used for the Tycho-2 proper motions. Thus, within the sky area covered, UCAC2 supersedes the Tycho-2 astrometry for stars 10th magnitude and fainter, providing the most precise positions and proper motions available today for catalogs of comparable area coverage. Users should note some UCAC limitations. Stars brighter than about R = 10, and in particular those brighter than R = 9, can suffer from overexposure effects and generally are based on 2 images of short exposures only. Their errors are higher and this is reflected in the catalog; they should be used with caution when the strictest astrometry is required. The UCAC observations provide only crude magnitudes in a single, non-standard bandpass (between V and R). To make the catalog more useful to the astronomy community, the Two Micron All Sky Survey \\citep{2mass} J, H, and K$_{S}$ infrared magnitudes are included for the matched sources (99.5\\% of the total UCAC sources). UCAC2 does not provide any trigonometric parallaxes. Systematic errors in the UCAC2 positions are 5 to 10 mas; although very small, these are larger than in the Hipparcos Catalogue. Along with UCAC2 superseding UCAC1, users should note that the UCAC1 was an observational catalog with attached, preliminary proper motions. UCAC2 is a compiled catalog of positions and proper motions referred to a standard epoch (J2000.0); the mean CCD observational position is not published. The level of completeness (about 80\\%) is the same for UCAC1 and UCAC2, avoiding all ``problem cases\" such as elongated images and blended images of close double stars. For the final release (UCAC3) likely both the mean observational data and the ``best\" compiled positions and proper motions will be published, with major improvements in completeness. ", "conclusions": "UCAC2 provides the most accurate positions and proper motions available today for most of the stars in the 9th to 16th magnitude range and the 86\\% of the sky covered so far. External, random errors are close to the quoted, internal errors. The 2MASS infrared photometry added to the data will be of benefit for the user, particularly for stellar statistics and galactic kinematics investigations. The average error of proper motions for the R = 13 to 16 magnitude stars is about 6 mas/yr, dramatically improved over the UCAC1, thanks to the inclusion of the Yellow Sky catalog. For brighter stars, with the inclusion of Hipparcos, Tycho, AGK2 and all catalogs used for the Tycho-2 construction, proper motion errors in the 1 to 2 mas/yr range could be achieved. The high positional accuracy of UCAC at current epoch has been exploited in the minor planet community and has proven essential for occultation predictions \\citep{mpocc} \\footnote{see also \\url{http://www.asteroidoccultation.com} and \\url{http://mpocc.astro.cz}}. The goal, providing a densification of reference stars beyond the Hipparcos / Tycho-2 catalogs, has been achieved. The average density of UCAC2 is 1,360 stars per square degree, with a positional accuracy of stars to 14th magnitude close to the current epoch position errors of the Hipparcos Catalogue. For most applications UCAC2 supersedes even Tycho-2 in the sky area covered for stars fainter than about 9th magnitude. However, UCAC2 is limited by remaining systematic errors on the 5 to 10 mas level, which -- although very small -- is significantly worse than for the Hipparcos Catalogue. Systematic errors in the UCAC2 proper motions have not been investigated in great detail yet. Comparisons with identified very distant sources indicate no obvious problem, with expected systematic errors on the 1 mas/yr level. Using UCAC2 for determining a possible system offset (rotation) between HCRF and ICRF at current epochs via observations of counterparts of extragalactic radio sources is severely limited by the remaining systematic errors in UCAC2 positions. To overcome this, as part of the UCAC project, these sources are observed with deep CCD images. The same fields are simultaneously observed with the UCAC astrograph on the east and west of the pier. These special observations are not included in the UCAC2 release. Mean positions derived from these additional observations will have much smaller systematic errors. One more final data release of UCAC is planned after the completion of the all-sky survey. It is envisioned that the pixel data will be re-processed for UCAC3, which should slightly improve the astrometric accuracy. The main advantage will be a completeness level to over 99\\%, thus providing accurate positions for many known and new double stars. The astrograph could be used for future projects \\citep{smtel}; however, a bigger telescope is required to make significant progress in further densification efforts. USNO has plans for a dedicated, astrometric, robotic, wide-field telescope for an all-sky survey to about 20th magnitude with positions on the 5 to 10 mas level to R = 18$^{m}$ \\citep{urat}." }, "0403/astro-ph0403132_arXiv.txt": { "abstract": "We present the discovery of a circumstellar dust disk surrounding AU Microscopium (AU Mic, GJ 803, HD 197481). This young M star at 10 parsec has the same age and origin as $\\beta$ Pictoris, another nearby star surrounded by a dust disk. The AU Mic disk is detected between 50 AU and 210 AU radius, a region where dust lifetimes exceed the present stellar age. Thus, AU Mic is the nearest star where we directly observe the solid material required for planet formation. Since 85\\% of stars are M-type, the AU Mic disk provides new clues on how the majority of planetary systems might form and evolve. ", "introduction": "About 15\\% of nearby main-sequence stars exhibit an excess of far-infrared radiation that points to the existence of circumstellar dust grains ({\\it1}). Dust grains have short lifetimes, and their continued presence implies a source of replenishment. The solar system has a disk-like distribution of dust that is continually replenished by the sublimation of comets and collisions between asteroids. We therefore infer that similar populations of undetected parent bodies produce dusty debris disks around infrared excess stars. Moreover, evidence for planets can be found by matching density variations in debris disks to theoretical models of how planets gravitationally perturb these disks ({\\it2}, {\\it3}, {\\it4}). In effect, circumstellar debris disks are a signpost for the existence of extrasolar planetary systems. Direct images of debris disks are rare. Starlight reflecting off optically thin debris disks is detected at optical and near-infrared wavelengths in only three cases - $\\beta$ Pictoris, HR 4796A and HD 141569 ({\\it5}, {\\it6}, {\\it7}). In three more cases - Vega, Fomalhaut, and $\\epsilon$ Eridani - debris disk structure is seen only at thermal infrared wavelengths ({\\it8}, {\\it9}). $\\beta$ Pic, HR 4796A and HD 141569 are relatively young ($<$20 Myr) main sequence stars that have the largest disk masses, and hence represent the more detectable of the debris disk systems. $\\beta$ Pic has Galactic space motions in common with two M stars, AU Mic and AT Mic, that have ages $\\sim$20 Myr ({\\it10}). These stars and $\\beta$ Pic may be co-eval sister stars that have separated in space over time due to the small differences in their space motions. A total of 17 stars may be members of this group with age 8$-$20 Myr ({\\it11}). AU Mic has significant infrared excess at 60 $\\mu$m ({\\it12}, {\\it13}) and recent sub-millimeter data reveal the the presence of cold (40 K) dust, a total dust mass roughly three times smaller than that of $\\beta$ Pic (Table 1), an absence of molecular gas, and a lack of grains within 17 AU of the star ({\\it14}). ", "conclusions": "The existence of morphologically similar dust disks around AU Mic and $\\beta$ Pic supports the hypothesis that these are sister stars born at the same time and location. However, the two disks are not twins. The total mass of dust estimated from the spectral energy distributions is 3.3 times greater for $\\beta$ Pic relative to AU Mic (Table 1). The relative brightnesses of the two disks in optical data are consistent with this result. To make the comparison, we imagine placing the $\\beta$ Pic dust disk around AU Mic. In Figure 2, we include the midplane surface brightness profile for $\\beta$ Pic using data from ({\\it16}) that is now scaled by factors that account for the AU Mic heliocentric distance and stellar luminosity. We find that if the disk of $\\beta$ Pic surrounded AU Mic it would be about 1.5 mag arcsec$^{-2}$ brighter than what we measure for the AU Mic disk (Fig. 2). This corresponds to a factor of four greater scattering cross section of $\\beta$ Pic grains relative to AU Mic grains. If we assume that the two disks have exactly the same structure, grain properties, and viewing geometry then the AU Mic disk requires a dust mass that is four times smaller than that of $\\beta$ Pic. Future observations of disk properties such as the inclination of AU Mic will elucidate the validity of these assumptions, but this result is consistent with the infrared dust luminosity. The underlying grain properties are also likely to differ due to the weak radiation environment of an M star relative to an A star. AU Mic is 3.6 times less massive than $\\beta$ Pic, and 87 times less luminous (Table 1). For the AU Mic disk, the collision timescale at 100 AU radius is 0.2$-$1.8 Myr assuming a dust optical depth of $\\tau\\sim$10$^{-3} -10^{-4}$, respectively (Fig. S4). At 200 AU, near the outer boundary of the detected disk, the collision timescale is 0.5$-$5.0 Myr. Given an age of 8$-$20 Myr for AU Mic, most disk particles have undergone at least one collision. However, as objects are shattered into smaller pieces, the radiation pressure force around AU Mic is too weak to remove the fragments ({\\it19}). They can be removed by the system either by joining together to form larger objects, or by spiraling into the star by Poynting-Robertson (PR) drag. The PR timescales at 100 AU are 0.2$-$1.8 Gyr for 1$-$10 $\\mu$m particles, respectively $-$ many times longer than the age of the system ({\\it1}). For $\\beta$ Pic, on the other hand, grains a few microns and smaller are quickly ejected by radiation pressure and the disk mass diminishes over time ({\\it20}). The AU Mic disk should preserve a larger population of sub-micron sized grains, and the mass of solid objects observed today should approximate the primordial disk mass. In other words, most of the disk seen in our optical scattered light image may consist of primordial solid material. Within $\\sim$50 AU of the star, the timescales for grain removal by collisions and PR drag become significantly shorter than the stellar age. Primordial dust at the inner limit of our images (Fig. 1, 2) has mostly vanished, and the grains observed here, as well as those discovered as close as 17 AU from the star ({\\it14}), must be continually replenished by the collisional erosion of much larger objects such as comets and asteroids. The existence of planetesimals in this region lends plausibility to the argument that the same objects will form planets by accretion. Given that AU Mic is only $\\sim$10 Myr old, we may be able to observe planets that are still in the process of accreting mass, or at least discern disk structure that is sculpted by planet-mass bodies. Because AU Mic is closer to the Sun than $\\beta$ Pic, the 2$-$30 AU zone where terrestrial and gas giant planets might form can be resolved by current and future instrumentation (Fig. S5). Planets around AU Mic may also be detected by indirect methods. The low stellar mass means that the star will display a significant astrometric reflex motion (2 milli-arcsec for a Jupiter-analog). The near edge-on orientation favors planet detection by transits of the stellar photosphere. Finally, if a planet is detected by radial velocity techniques, then the near edge-on orientation gives the planet mass by constraining the sin($i$) ambiguity intrinsic to these measurements. \\newpage \\noindent{\\bf References and Notes} \\noindent1. D. E. Backman, F. Paresce, in {\\it Protostars and Protoplanets III}, E. H. Levy and J. I. Lunine, Eds. (University of Arizona Press, Tucson, 1993), pp. 1253-1304. \\noindent2. F. Roques, H. Scholl, B. Sicardy, B. A. Smith, $Icarus$ {\\bf 108}, 37 (1994). \\noindent3. J. -C. Liou, H. A. Zook, {\\it Astron. J.} {\\bf 118}, 580 (1999). \\noindent4. L. M. Ozernoy, N. N. Gorkavyi, J. C. Mather, T. A. Taidakova {\\it Astrophys. J.} {\\bf 537}, L147 (2000). \\noindent5. B. A. Smith, R. J.Terrile, $Science$ {\\bf 226}, 1421 (1984). \\noindent6. G. Schneider et al., {\\it Astrophys. J.} {\\bf 513}, L127 (1999). \\noindent7. M. D. Silverstone et al., {\\it Bull. Am. Astron. Soc} {\\bf 30}, 1363 (1998). \\noindent8. W. S. Holland et al., $Nature$ {\\bf 392}, 788 (1998). \\noindent9. J. S. Greaves et al., {\\it Astrophys. J.} {\\bf 506}, L133 (1998). \\noindent10. D. Barrado y Navascues, J. R. Stauffer, I. Song, J-. P. Caillault, {\\it Astrophys. J.} {\\bf 520}, L123 (1999). \\noindent11. B. Zuckerman, I. Song, M. S. Bessell, R. A. Webb, {\\it Astrophys. J.} {\\bf 562}, L87 (2001). \\noindent12. V. Tsikoudi, {\\it Astron. J.} {\\bf 95}, 1797 (1988). \\noindent13. I. Song, A. J. Weinberger, E. E. Becklin, B. Zuckerman, C. Chen, {\\it Astron. J.} {\\bf 124}, 514 (2002). \\noindent14. M. C. Liu, B. C. Matthews, J. P. Williams, P. G. Kalas, {\\it Astrophys. J.}, in press (2004). \\noindent 15. Materials and methods are available as supporting material on {\\it Science Online}. \\noindent16. P. Kalas, D. Jewitt, {\\it Astron. J.} {\\bf 110}, 794 (1995). \\noindent17. To increase the signal-to-noise of the data shown in Fig. 1, we binned the data 3$\\times$3 pixels and then smoothed by a Gaussian function with $\\sigma$=0.5 pixel. This smoothed image was used to find the maximum outer extent of the disk. All other measurements were made using the unbinned and unsmoothed image shown in Fig. 1. \\noindent18. P. Kalas, D. Jewitt, {\\it Astron. J.} {\\bf 111}, 1347 (1996). \\noindent19. R. Saija, et al. {\\it Mon. Not. R. Astron. Soc.} {\\bf 341}, 1239 (2003). \\noindent20. P. Artymowicz, {\\it Astrophys. J.} {\\bf 335}, L79 (1988). \\noindent21. M. S. Bessel, F. Castelli, B. Plez, {\\it Astron. Astrophys.} {\\bf 333}, 231(1998). \\noindent22. F. Crifo, A. Vidal-Madjar, R. Lallement, R. Ferlet, M. Gerbaldi, {\\it Astron. Astrophys.} {\\bf 333}, L29 (1997). \\noindent23. J. D. Larwood, P. G. Kalas, {\\it Mon. Not. R. Astron. Soc.} {\\bf 323}, 402 (2001). \\noindent24. W. R. F. Dent, H. J. Walker, W. S. Holland, J. S. Greaves, {\\it Mon. Not. R. Astron. Soc.} {\\bf 314}, 702 (2000). \\noindent25. This work has been supported by the NASA Origins Program under grant NAG5-11769, and the NSF Center for Adaptive Optics, managed by the University of California at Santa Cruz under cooperative agreement No. AST-9876783. BCM acknowledges support from NSF grant \\#0228963. MCL acknowledges support from a Hubble Postdoctoral Fellowship (NASA Grant HST-HF-01152.01). The authors acknowledge the insightful contributions of two anonymous referees. \\noindent {\\bf Supporting Online Material}\\\\ www.sciencemag.org\\\\ Materials and Methods\\\\ Figs. S1, S2, S3, S4 and S5\\\\ Table S1\\\\ \\newpage \\begin{figure} \\epsscale{0.8} \\plotone{fig1.ps} \\caption{ The disk surrounding AU Mic seen in optical scattered light. North is up, east is left, and each side of this false-color image corresponds to 60$\\arcsec$. The central dark region is produced by the 9.5$\\arcsec$ diameter focal plane occulting spot which is suspended by four wires and completely masks our direct view of the star. This image represents 900 seconds total integration in the $R$ band and each pixel corresponds to 4 AU at the distance to AU Mic. Residual light evident near the occulting spot edge in the NE-SW direction is attributed to asymmetries in the point-spread function caused by instrumental scattering and atmospheric seeing. \\label{fig1}} \\end{figure} \\begin{figure} \\epsscale{0.8} \\plotone{fig2.ps} \\caption{\\small Midplane surface brightness as a function of radius. The midplane was sampled between 5$\\arcsec$ and 21$\\arcsec$ radius along a strip 1.2$\\arcsec$ wide. We show the mean value from two nights of data with two different PSF subtraction techniques. The error bars represent one standard deviation of a single measurement. We fit the data between 6$\\arcsec$ and 16$\\arcsec$ radius with power laws that give indices -3.6 and -3.9 for the NW and SE midplanes of AU Mic, respectively. The radial profile for the NW midplane has a significant brightness enhancement at $\\sim$9$\\arcsec$ radius that is either intrinsic to the disk or a background object. We also plot the surface brightness of the $\\beta$ Pic disk from ({\\it16}), but with the surface brightness uniformly 3.0 mag arcsec$^{-2}$ fainter to simulate the existence of $\\beta$ Pic's disk around AU Mic at 9.9 pc. This scaling takes into account the fact that the absolute $R$-band magnitude of AU Mic is 5.6 mag fainter than $\\beta$ Pic (Table 1), and at a constant angular radius the $\\beta$ Pic disk is roughly a factor of ($d_{AU Mic} / d_{\\beta Pic})^{-3.6}$ = (9.9 / 19.3)$^{-3.6}$ =11.3 times brighter (i.e. 2.6 mag brighter; see Eqn. 4 in {\\it18}). \\label{fig2} } \\end{figure} \\clearpage \\begin{deluxetable}{lllll} \\tabletypesize{\\small} \\tablecaption{Star (rows 1-8) and disk (rows 9-13) properties for AU Mic and $\\beta$ Pic. The stellar parameters for AU Mic are derived from data given by ({\\it10}, {\\it13}, {\\it21}). For $\\beta$ Pic's stellar parameters, we use ({\\it22}) and references therein. The $\\beta$ Pic disk $R$-band surface brightness (SB) at 6$\\arcsec$ radius is given in ({\\it16}), while its maximum extent is given in ({\\it23}). \\label{tbl-1}} \\tablewidth{0pt} \\tablehead{ \\colhead{} & \\colhead{AU Mic} & \\colhead{$\\beta$ Pic} } \\startdata Spectral Type \t\t\t\t\t\t&M1Ve \t\t& A5V \\\\ Mass (M$_\\odot$)\t\t\t\t\t&0.5\t\t\t&1.8\\\\ T$_{eff}$ (K) \t\t\t\t\t\t&3500 \t\t& 8200\\\\ Luminosity (L$_\\odot$)\t\t\t\t&0.1\t\t\t&8.7\\\\ Distance (pc) \t\t\t\t\t\t&9.9\t\t\t& 19.3 \\\\ $V$ (mag)\t\t\t\t\t\t&8.8\t\t& 3.9 \\\\ M$_V$ (mag)\t\t\t\t\t\t&8.8\t\t& 2.4 \\\\ $V-R$\t\t\t\t\t\t\t&0.88\t\t& 0.08\\\\ Disk SB (6$\\arcsec$)\t\t\t\t&20.1$\\pm$0.3\t&15.4$\\pm$0.3\\\\ SB fall-off\\tablenotemark{a}\t\t\t&-3.6 to -3.9\t&-3.8 to -4.1 \\\\ Max. Radius (AU)\t\t\t\t\t&210\t\t&1835 \\\\ $\\tau$=$L_{disk}$/$L_{bol}$\\tablenotemark{b}\t&6.1$\\times$10$^{-4}$ & 3$\\times$10$^{-3}$ \\\\ Total Dust Mass (g)\\tablenotemark{c}\t&6.6$\\times$10$^{25}$& 2.2$\\times$10$^{26}$ \\\\ \\enddata \\tablenotetext{a}{Value of exponent for a power-law fit to disk $R$-band surface brightness fall-off between 6-16$\\arcsec$ radius. The shallower surface brightness profiles correspond to the NW and NE brightness profiles of AU Mic and $\\beta$ Pic, respectively. The $\\beta$ Pic values are taken from ({\\it16}). } \\tablenotetext{b}{The fractional dust luminosity, assuming an optically thin disk, determined by taking the ratio of excess infrared luminosity to stellar bolometric luminosity. The values are obtained from ({\\it14}) and ({\\it1}) for AU Mic and $\\beta$ Pic, respectively. } \\tablenotetext{c}{Dust mass from model fits to the spectral energy distributions taken from ({\\it14}) for AU Mic, and ({\\it24}) for $\\beta$ Pic. } \\end{deluxetable} \\clearpage" }, "0403/astro-ph0403418_arXiv.txt": { "abstract": "We present imaging observations of the evolved star IRC+10216 in the CS J=14--13 line at 685.4 GHz and associated submillimeter continuum at $\\sim 2''$ resolution made with the partially constructed Submillimeter Array. The CS J=14--13 line emission from the stellar envelope is well resolved both spatially and spectrally. The strong central concentration of the line emission provides direct evidence that CS is a parent molecule that forms close to the stellar photosphere, in accord with previous images of the lower excitation CS J=2--1 line and inferences from unresolved observations of vibrationally excited transitions. The continuum emission is dominated by a compact, unresolved component, consistent with the photospheric emission, that accounts for $\\sim20\\%$ of the broadband 450~$\\mu$m flux. These are the first interferometer imaging observations made in the semi-transparent 450~$\\mu$m atmospheric window. ", "introduction": "The nearby carbon star IRC+10216 (CW Leo, IRAS 09452+1330) is surrounded by a dusty, expanding envelope with relatively simple geometry and kinematics that has been the subject of many studies of astrochemical processes. A rich chemistry develops in the envelope as gas is lost from the star at a high rate ($\\sim3\\times10^{-5}$~M$_{\\odot}$~yr$^{-1}$, Crosas \\& Menten 1997) and is modified by thermodynamic equilibrium and non-equilibrium reactions, photochemical reactions, ion-molecule reactions, and the condensation of dust grains. The range of physical conditions in the envelope make it especially well suited to observations of rotational emission lines from molecules, and numerous species have been identified in spectral scans at millimeter and submillimeter wavelengths (e.g. Cernicharo, Guelin \\& Kahane 2000; Groesbeck, Phillips \\& Blake 1994, Avery et al. 1992). The close proximity ($\\sim150$~pc) and copious mass loss of IRC+10216 allows studies of the circumstellar envelope on a wide range of size scales. On the largest scales, an extensive envelope of dust (Mauron \\& Huggins 1999) and gas (Fong, Meixner \\& Shah 2003) has been found to extend to $\\sim0.15$~pc ($\\sim3\\farcm3$) in radius. On much smaller scales, interferometric imaging at millimeter wavelengths has revealed the spatial distribution of molecular species within the central arcminute (e.g. Bieging \\& Tafalla 1993, Guelin, Lucas \\& Cernicharo 1993, Lucas \\& Guelin 1999). Parent molecules are found to be centrally peaked (e.g. SiS, HCN), while daughter molecules are found with shell distributions (e.g. CN, C$_2$H, HNC). High angular resolution is especially important for imaging lines of high excitation that emerge from the innermost regions of the envelope, where the temperatures and densities are highest. Since the high excitation rotational lines of many common molecules lie at short submillimeter wavelengths, the Submillimeter Array\\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.} (SMA) opens new possibilities for probing physical conditions and chemistry by enabling imaging observations with arcsecond resolution. In this {\\em Letter}, we present observations of IRC+10216 in CS J=14--13 emission made with the partially constructed SMA (Ho, Moran \\& Lo 2004). These are the first interferometer images from observations in the atmospheric window centered near 450~$\\mu$m, where $\\sim$30\\% or better transmission occurs at high, dry sites like the summit of Mauna Kea a significant fraction of the time. They build on the pioneering efforts of Carlstrom et al. (1994), who obtained fringes on the nearby single baseline between the Caltech~Submillimeter~Observatory and the James~Clerk~Maxwell~Telescope. The results provide constraints on sulfur chemistry in the inner wind of the IRC+10216 prototype carbon-rich circumstellar envelope. ", "conclusions": "\\subsection{Spectrum at Image Center} Figure~\\ref{fig:cs_spectrum} shows the spectrum from the center of the image cube for the full USB correlator band, averaging groups of 8 channels, corresponding to a velocity width of 2.85 km~s$^{-1}$. The CS J=14--13 line is the prominent feature centered at $-26$~km~s$^{-1}$ and an approximate width at zero power of $\\sim30$~km~s$^{-1}$, values which are consistent with the systemic velocity of the source and the expansion velocity of the envelope, respectively. The line shape is difficult to interpret because spatial filtering by the interferometer results in missing flux that varies with velocity across the profile (see \\S\\ref{discussion:line}). The peak brightness temperature of the line at the full angular resolution of the data is well over 100~K. An additional feature appears in the spectrum near $-100$~km~s$^{-1}$ with a double peaked line shape and an approximate width consistent with the circumstellar expansion. A search of the JPL spectral line catalog suggests a tentative identification of this feature with the C$_3$H$_2$ $8(4,4)-7(3,5)$ line at 685.6125565~GHz. This species was previously detected in transitions at lower frequencies (Kawaguchi et al. 1995, Cernicharo et al. 2000). No alternative candidates line were found in the catalog of Cernicharo et al. (2000) that contains 1050 molecular species (Cernicharo, private communication). If the assignment is correct, then the line might be thought to emerge from the external shell where various carbon chain radicals are found (Guelin et al. 1993). This feature appears to be spatially compact, though, which perhaps renders the identification problematic. \\subsection{Continuum Emission} Figure~\\ref{fig:cont} shows the continuum image obtained from the channels free of strong line emission. Visibilities from both sidebands, separated by 10~GHz, were combined in a multifrequency synthesis, resulting in an effective frequency of 680 GHz. The continuum peak position is consistent with previous determinations, in particular the 95~GHz observation from the IRAM PdBI (Guelin et al. 1993), which is marked by the cross. The peak continuum flux is 3.8~Jy, with an uncertainty dominated by systematic effects. Previous bolometer observations of IRC+10216 at 450~$\\mu$m indicate that the SMA recovers approximately 20\\% of the broadband flux at this wavelength. Sandell (1994) noted that IRC+10216 has long period variations in the submillimeter and found $19\\pm3$~Jy in an $18''$ aperture near the 635 day maximum. Jenness et al. (2002) analyzed several years of SCUBA observations of IRC+10216 at 850~$\\mu$m and 450~$\\mu$m and found a period and maximum flux consistent with the earlier single pixel measurements. The SMA observations were made approximately a month before the predicted maximum, when $\\sim95\\%$ of the peak flux would be expected. Remarkably, the CS J=14--13 line emission {\\em alone} accounts for $1\\%$ of the flux in the $\\sim68$~GHz wide filter used for the 450~$\\mu$m bolometer measurements. Given the many strong spectral lines excited in the inner envelope at these high frequencies, it is clear that spectral line emission makes an important contribution to the broadband ``continuum'' flux. A large fraction of the continuum emission from the compact component detected by the SMA at 680~GHz likely comes from the stellar photosphere. Lucas \\& Guelin (1999) report a ``point source'' component in IRAM PdBI observations of $65\\pm1$~mJy at 89~GHz and $486\\pm7$~mJy at 242~GHz, identified as photospheric emission. Assuming an optically thick spectrum, $S_\\nu\\propto\\nu^2$, an extrapolation to 680~GHz gives 3.8~Jy, consistent with the measured value. Thermal emission from dust in the inner envelope cannot contribute substantial additional flux in the small synthesized beam of the SMA at this high frequency. \\subsection{CS J=14--13 Emission} \\label{discussion:line} Figure~\\ref{fig:cs_channels} shows images for a series of velocity intervals that span the CS J=14--13 line. Below each of the images is a plot of visibility amplitude vs. $(u,v)$ distance for the corresponding velocity interval to give an idea of spatial extent, which is not easy to ascertain from the images. The overall structure in the visibility amplitudes is consistent with simple expectations for a spherically expanding envelope. At the extreme velocities, the amplitude is approximately constant, as expected for the unresolved ``caps'' of blueshifted and redshifted emission, while at intermediate velocities, the fall-off of visibility amplitude with $(u,v)$ distance demonstrates that the emission is spatially extended and resolved. If the gas is expanding radially with approximately spherical symmetry, then the narrow velocity interval around the central velocity corresponds to a cross section through the envelope in the plane of the sky. The central peak of CS J=14--13 emission distribution provides additional direct evidence that CS is ``parent'' molecule, in accord with the detection of ro-vibrational transitions in absorption in the infrared (Keady \\& Ridgeway 1993), the detection of vibrationally excited CS emission (Turner 1987, Lucas \\& Guelin 1999, Highberger et al. 2000), and interferometric imaging of CS J=2--1 emission (Lucas et al. 1995). The CS molecules must originate close to the stellar photosphere, and they are lost from the expanding envelope through reactions that build more complex species, perhaps aided by shocks as suggested by Willacy \\& Cherchneff (1998). The J=14-13 emission distribution shows no indication of the extended $\\sim15''$ radius ring visible in the J=2-1 line, but emission on that size scale, if present, could not be detected on account of the small SMA field of view. The reaction network for sulfur bearing carbon chains has been explored by Millar, Flores \\& Markwick (2001), who concluded that an initial CS abundance of $4\\times10^{-6}$ is needed to match observations of C$_3$S and C$_5$S. Models where CS is not a parent species, in which the CS fractional abundance rises from $<10^{-11}$ at radius $10^{16}$~cm, or $4\\farcs$5 at 150~pc, to a peak more than four orders of magnitude larger at radius $10^{17}$~cm, could not match observations of the longer carbon chains. The SMA observations provide a robust {\\em lower limit} to the CS abundance within a radius of $\\sim34$~R$_*$ ($2.25\\times10^{15}$~cm) that supports the parent species scenario. Under the assumption of a thermalized level population, the total CS column density is given by \\begin{equation} N(CS) = \\frac{3 k f}{8 \\pi^3 B \\mu^2} \\frac{\\exp{(hBJ_l(J_l+1)/kT)}}{J_l+1} \\frac{T + hB/3k}{1- \\exp{(-h\\nu/kT)}} \\int \\tau dv, \\end{equation} where the molecular constants are $\\mu=1.96$ Debye, $B=24.584$ GHz, and the filling factor $f$ may be estimated as (e.g. Scoville et al. 1986) \\begin{equation} f = \\frac{T_R^*/\\eta_c}{(h\\nu/k)/(\\exp{(h\\nu/kT)-1})(1 - \\exp{-\\tau}) }. \\end{equation} The lower limit to the line flux in a $\\sim2''$ beam is $\\sim 5000$~Jy~km~s$^{-1}$, or $\\sim 1250$~K~km~s$^{-1}$ in brightness units. Assuming small optical depth, $\\tau<<1$, a kinetic temperature equal to the excitation of the upper level of the observed transition (246.8~K), which minimizes the column density required to explain the observed line flux, and a spherical volume with molecular hydrogen density equal to the critical density of the transition of $3.2\\times10^8$~cm$^{-3}$, results in a lower limit to the CS fractional abundance of $\\sim3.4\\times10^{-9}$. The abundance close to the photosphere is likely substantially higher than this estimate; Highberger et al. (2000) derive $3-7\\times10^{-5}$ at a radius $\\sim14$~R$_*$ ($9\\times10^{14}$~cm) from their multi-transition single dish observations. Though the uncertainties are large, models such as those investigated by Millar et al. (2001) where the CS abundance rises from $<10^{-11}$ at radius $10^{16}$~cm are clearly in conflict with the high resolution images of the CS J=14--13 emission. The SMA observations show directly that the CS molecules are formed close to the stellar photosphere. Detailed radiative transfer models that account for the observed spatial distribution are needed for an accurate determination of the CS abundance in the inner envelope. The effect of pulsational shocks could be important in establishing the CS abundance as this species is injected into the wind, and careful modeling may provide insight and constraints on the shock chemistry. Imaging the CS J=14--13 line in the inner envelope of IRC+10216 offers a first demonstration of the efficacy of the SMA in the semi-transparent 450~$\\mu$m atmospheric window. Calibration at these high frequencies should become considerably easier as more receivers for these frequencies are deployed at the SMA, the correlator bandwidth is increased to 2~GHz, and simultaneous operation of a lower frequency band enables phase transfer. Many more evolved stars, as well as other objects detected but with presently unresolved emission at short submillimeter wavelengths, are now accessible to imaging at the arcsecond scale." }, "0403/astro-ph0403304_arXiv.txt": { "abstract": "We present the first results from the \\xmm Galactic Plane Survey (XGPS). In the first phase of the programme, 22 pointings were used to cover a region of approximately three square degrees between 19\\deg -- 22\\deg in Galactic longitude and $\\pm$0.6\\deg in latitude. In total we have resolved over 400 point X-ray sources, at $\\geq 5 \\sigma$ significance, down to a flux limit of $\\sim2 \\times 10^{-14}$ \\ergseccm (2--10~keV). The sources exhibit a very wide range of spectral hardness, with interstellar absorption identified as a major influence. The source populations detected in the soft (0.4--2 keV) band and hard (2--6 keV) band show surprisingly little overlap. The majority of the soft sources appear to be associated with relatively nearby stars with active stellar coronae, judging from their high coincidence with bright stellar counterparts. The combination of the XGPS measurements in the hard X-ray band with the results from earlier surveys carried out by \\asca and \\chan reveals the form of the low-latitude X-ray source counts over 4 decades of flux. It appears that extragalactic sources dominate below $\\sim10^{-13}$ \\ergseccm (2--10~keV), with a predominantly Galactic source population present above this flux threshold. The nature of the faint Galactic population observed by \\xmm remains uncertain, although cataclysmic variables and RS CVn systems may contribute substantially. \\xmm observes an enhanced surface brightness in the Galactic plane in the 2--6 keV band associated with Galactic Ridge X-ray Emission (GRXE). The integrated contribution of Galactic sources plus the breakthrough of extragalactic signal accounts for up to 20 per cent of the observed surface brightness. The XGPS results are consistent with the picture suggested from a deep \\chan observation in the Galactic plane, namely that the bulk of the GRXE is truly diffuse. ", "introduction": "With the current generation of X-ray astronomy missions, we are for the first time able to carry out high sensitivity, coherent surveys of selected regions of the Galactic plane. In particular, the \\xmm mirrors afford a large collecting area ($\\sim 4650 \\rm~cm^{-2}$ total geometric area) with good spatial resolution (FWHM $\\sim6''$ and HEW $\\sim15''$ on-axis) over a wide field of view ($30'$ diameter). In combination with the EPIC CCD cameras, this provides an excellent facility for surveying sky regions subtending many square degrees down to relatively faint flux levels in both the hard ($> 2$ keV) and soft ($< 2$ keV) X-ray bands. The goal of the \\xmm Galactic Plane Survey (XGPS) is two-fold. The first objective is to study the properties of the Galactic X-ray source population at intermediate flux levels (down to $\\sim 2 \\times 10^{-14}$ \\ergseccm~in the 2--10~keV band but an order of magnitude fainter in flux terms in the softer 0.4--2~keV band). The second is to search for extended, low X-ray surface brightness features including variations in the underlying diffuse Galactic Ridge X-ray Emission (GRXE; \\citeb{a255}; \\citeb{n317}; \\citeb{pasj38}; \\citeb{a404}; \\citeb{a491}; \\citeb{a505}). The nature of the X-ray source population at high X-ray fluxes was established by early all-sky surveys and subsequent identification programmes, which revealed that the brightest sources in our Galaxy are predominantly X-ray binaries and supernova remnants. More sensitive surveys of the Galactic plane have since been made, including those made by {\\it ROSAT} (\\citeb{aa246}) and {\\it ASCA} (\\citeb{a134}) complemented by the serendipitous surveys carried out with the Einstein observatory (\\citeb{a278}). Together, these surveys have provided some glimpses of the X-ray source population at lower X-ray fluxes, and hence effectively at lower X-ray luminosities for Galactic objects, although the picture is far from complete. At soft X-ray energies ($<2$ keV) {\\it ROSAT} studies in particular have shown that coronal emission from relatively nearby active stars dominates (\\eg {\\citeb{aa246b}). Above 2 keV the characteristics of the harder population are far less well-defined, although it is clear that accreting binary sources (both X-ray binaries and cataclysmic variables) make a significant contribution. To date the XGPS survey has been targeted at several locations in the Galactic segment between the Galactic Centre and the Scutum Spiral Arm. Here we report the results from the first phase of the XGPS (hereafter XGPS-I), which has entailed a total of 22 \\xmm pointings, covering a region of approximately three square degrees between 19\\deg -- 22\\deg in Galactic longitude and $\\pm$0.6\\deg in latitude. Over 400 discrete point-like X-ray sources have been detected in XGPS-I and in this paper we focus on the properties of this source population and the contribution these discrete sources make to the GRXE. In a second paper (Hands et al. 2004, in preparation) we will present the results of a search for low-surface brightness, spatially extended X-ray sources in the XGPS-I fields and also report on the properties of the underlying diffuse GRXE. \\begin{table*} \\begin{center} \\caption{Observation log for the XGPS-I. \\label{tab:log}} \\vspace{0.1in} \\begin{tabular}{lcllrrrr} \\hline Field & & \\multicolumn{2}{c}{Field Centre (J2000)} & MOS & pn & MOS & pn\\\\ Name & Observation Date & R.A. & Dec. & exposure$^a$ & exposure$^b$ & fraction$^c$ & fraction$^c$ \\\\ \\hline Ridge 1 & 2000-10-08 & 18 26 00.4 & -12 14 55.9 & 8393 & 5914 & 0.95 & 0.85 \\\\ Ridge 2 & 2002-09-21 & 18 26 48.4 & -11 52 48.7 & 13667 & 12046 & 1.00 & 1.00 \\\\ Ridge 3 & 2000-10-11 & 18 27 36.4 & -11 30 40.4 & 12044 & 9648 & 0.95 & 0.85 \\\\ Ridge 4 & 2000-10-12 & 18 28 17.0 & -11 07 53.0 & 9256 & 12998 & 0.94 & 0.42 \\\\ Ridge 5 & 2002-09-17 & 18 29 06.0 & -10 45 03.0 & 13667 & 12046 & 0.98 & 0.93 \\\\ XGPS 1 & 2001-03-08 & 18 25 04.6 & -11 50 00.4 & 7794 & 4797 & 1.00 & 1.00 \\\\ XGPS 2 & 2001-03-10 & 18 27 34.0 & -12 09 20.0 & 9144 & - & 1.00 & - \\\\ XGPS 3 & 2001-03-10 & 18 25 49.0 & -11 28 42.7 & 9144 & - & 1.00 & - \\\\ XGPS 4 & 2001-03-10 & 18 28 19.4 & -11 48 05.2 & 9144 & - & 1.00 & - \\\\ XGPS 5 & 2001-03-22 & 18 26 35.7 & -11 07 32.1 & 9994 & 7348 & 0.96 & 0.95 \\\\ XGPS 6 & 2001-03-22 & 18 29 05.9 & -11 26 57.4 & 8794 & 6148 & 0.44 & 0.23 \\\\ XGPS 7 & 2001-03-24 & 18 27 21.2 & -10 46 17.8 & 7637 & 4998 & 0.95 & 0.98 \\\\ XGPS 8 & 2002-09-29 & 18 29 50.8 & -11 05 41.2 & 13167 & 11546 & 1.00 & 1.00 \\\\ XGPS 9 & 2001-03-26 & 18 28 06.4 & -10 25 10.0 & 11894 & 9248 & 0.92 & 0.62 \\\\ XGPS 10 & 2001-10-03 & 18 30 36.2 & -10 44 29.3 & 9767 & 8146 & 0.97 & 0.83 \\\\ XGPS 11 & 2002-09-19 & 18 29 43.9 & -10 24 09.6 & 8409 & 6788 & 0.92 & 0.81 \\\\ XGPS 12 & 2002-09-27 & 18 28 59.1 & -10 06 50.9 & 7367 & 5746 & 1.00 & 1.00 \\\\ XGPS 13 & 2003-04-10 & 18 31 25.5 & -10 24 32.5 & 6666 & 5047 & 0.73 & 0.66 \\\\ XGPS 14 & 2002-03-11 & 18 30 29.3 & -10 02 47.1 & 7269 & 4998 & 1.00 & 1.00 \\\\ XGPS 15 & 2002-03-27 & 18 29 36.6 & -~9 42 41.1 & 7274 & 4998 & 0.82 & 0.96 \\\\ XGPS 16 & 2002-03-15 & 18 32 06.5 & -10 01 51.8 & 7762 & 5486 & 1.00 & 1.00 \\\\ XGPS 17 & 2002-03-29 & 18 31 13.9 & -~9 41 39.4 & 7274 & 4998 & 0.99 & 0.94 \\\\ \\hline & & & & & \\\\ \\multicolumn{4}{l}{$^a$ Total exposure for the MOS 1 camera (s)} \\\\ \\multicolumn{4}{l}{$^b$ Total exposure for the pn camera (s)} \\\\ \\multicolumn{4}{l}{$^c$ Fraction of the exposure time used in producing images} \\\\ \\end{tabular} \\end{center} \\end{table*} ", "conclusions": "The XGPS-I survey, which covers approximately three square degrees of the Galactic Plane near $l=20$\\deg, has resulted in a catalogue containing over 400 discrete X-ray sources. The measured X-ray source counts trace the source population down to a limiting flux of $\\sim 2 \\times 10^{-14}$ \\ergseccm~in the 2--10 keV band at which point the source density is between 100--200 sources per square degree. Consistent with an earlier \\chan study, the source counts at this flux are predominately due to extragalactic sources, despite the fact that the fluxes of extragalactic objects are significantly suppressed by absorption in the Galactic plane. However, the conclusion of the present work is that at fluxes above $10^{-13}$ \\ergseccm (2--10 keV) Galactic source populations do come to the fore. The Galactic source population observed between $10^{-13}$ and $10^{-12}$ \\ergseccm~could comprise largely CVs and RS CVn systems with X-ray luminosities in the range $10^{30-32}$ \\ergsec~but the details remain uncertain on the basis of the X-ray information alone. Extensive programmes to identify and characterise optical/infra-red counterparts are required, although this will be taxing given the high obscuration and high object density in the Galactic plane. The present work demonstrates that the strategy of the XGPS programme, namely the use of shallow observations to give relatively wide angle coverage is close to optimum in terms of maximising the number of Galactic source detections." }, "0403/astro-ph0403074_arXiv.txt": { "abstract": "{ We have designed and tested a new relativistic Lagrangian hydrodynamics code, which treats gravity in the conformally flat approximation to general relativity. We have tested the resulting code extensively, finding that it performs well for calculations of equilibrium single-star models, collapsing relativistic dust clouds, and quasi-circular orbits of equilibrium solutions. By adding in a radiation reaction treatment, we compute the full evolution of a coalescing binary neutron star system. We find that the amount of mass ejected from the system, much less than a percent, is greatly reduced by the inclusion of relativistic gravitation. The gravity wave energy spectrum shows a clear divergence away from the Newtonian point-mass form, consistent with the form derived from relativistic quasi-equilibrium fluid sequences.} ", "introduction": "It has long been recognized that coalescing binary neutron star (NS) systems are a leading candidate to be the first observed source of gravity waves (GW). With LIGO, GEO, and TAMA all taking scientific data, and VIRGO in the commissioning stage, it is growing increasingly important to have quantitatively accurate predictions of the GW signals we expect to measure during the merger process. Besides their use in aiding detections, these predictions are crucial for determining important physical information about the mass, radius, and equation of state (EOS) of NS from GW observations. Calculations of binary NS coalescence have been performed for many years, beginning with studies in Newtonian gravity. It was recognized all along, however, that general relativity (GR) will play an important role during the merger, since the characteristic gravitational fields and velocities are squarely within the relativistic regime. As a result, increasingly sophisticated gravitational formalisms have been used in hydrodynamical calculations, starting with post-Newtonian treatments \\cite{FR1,FR2,FR3,APODR,SNO}, many of which were based on a formalism developed by Blanchet et al. \\cite{BDS} which includes all lowest-order 1PN effects as well as lowest order dissipative effects from gravitational radiation reaction losses. More recently, calculations have been performed in full general relativity \\cite{SU1,SU2,STU}. Unfortunately, the PN approximation breaks down during the merger when higher-order relativistic effects grow significant, and fully relativistic calculations typically introduce numerical instabilities which limit the amount of time for which a calculation will remain accurate. A middle ground is provided by the conformally flat (CF) approximation, developed originally by Wilson et al. \\cite{WMM}, which includes much of the non-linearity inherent in GR, but results in a set of coupled, non-linear, elliptic field equations, which can be evolved stably. We assume that the spatial part of the GR metric is equal to the flat-space form, multiplied by a conformal factor which varies with space and time, the metric taking the form \\begin{equation} ds^2=-(N^2-B_i B^i)dt^2-2B_i dt dx^i+A^2 \\delta_{ij}dx^i dx^j. \\end{equation} While this approach cannot reproduce the exact GR solution for a general matter configuration, it is exact for spherically symmetric systems, and yields solutions which agree with those calculated using full GR to within a few percent for many systems of interest \\cite{MMW}. ", "conclusions": "" }, "0403/astro-ph0403597_arXiv.txt": { "abstract": "We analyse the anisotropy of the cosmic microwave background (CMB) in hyperbolic universes possessing a non-trivial topology with a fundamental cell having an infinitely long horn. The aim of this paper is twofold. On the one hand, we show that the horned topology does not lead to a flat spot in the CMB sky maps in the direction of the horn as stated in the literature. On the other, we demonstrate that a horned topology having a finite volume could explain the suppression of the lower multipoles in the CMB anisotropy as observed by COBE and WMAP. ", "introduction": "A fundamental problem in cosmology is the large-scale geometry of the Universe, in particular its spatial curvature and topology. Since the Einstein gravitational field equations are differential equations, they determine the local properties of space-time, but not the global structure of the Universe at large. In the so-called concordance model of cosmology it is assumed that the Universe is at large scales spatially flat and possesses the trivial topology, implying that it has infinite volume. In the framework of inflationary scenarios, these properties are supposed to be determined by the initial conditions at the Big Bang. However, since we are lacking a theory of quantum gravity, the initial conditions cannot be derived from first principles. Instead, one can analyse the recent astronomical data, in particular on the cosmic microwave background (CMB) radiation, in order to deduce the curvature and the topology of the Universe at large scales. In 1992, COBE \\cite{Smoot_et_al_1992} made the spectacular discovery of the temperature fluctuations of the CMB and thus provided important clues about the early Universe and its time evolution. Expanding the observed temperature fluctuations $\\delta T(\\hat n)$ across the microwave sky into spherical harmonics $Y_l^m(\\hat n)$, yields the expansion coefficients $a_{lm}$ which in turn lead to the {\\it multipole moments} \\begin{equation} \\label{Eq:C_l} C_l \\; := \\; \\frac 1{2l+1} \\, \\sum_{m=-l}^l \\, | a_{lm} |^2 \\end{equation} and the {\\it angular power spectrum} $\\delta T_l^2 := l(l+1) C_l /(2\\pi)$. In particular, COBE \\cite{Smoot_et_al_1992} detected in the angular power spectrum of the CMB a low quadrupole moment $C_2$ corresponding to a strange suppression of power on large angular scales. It was soon realized that although the standard cosmological models are in agreement with the CMB anisotropy on small and medium scales, they fail to match the loss of power on large angular scales, especially those corresponding to the quadrupole moment. This observation was one of the motivations to study non-trivial topologies (see the reviews \\cite{Lachieze-Rey_Luminet_1995,Levin_2002}). In particular, compact hyperbolic universes were studied by several authors \\cite{Bond_Pogosyan_Souradeep_1998,Bond_Pogosyan_Souradeep_1999a,% Bond_Pogosyan_Souradeep_1999b,Cornish_Spergel_Starkman_1998,Aurich_1999,% Cornish_Spergel_1999,Inoue_Tomita_Sugiyama_1999,Aurich_Steiner_2000}. For these compact hyperbolic universes, it was shown that the non-trivial topology leads indeed to a suppression of $C_l$ for small values of $l$. Thus the observed loss of power on large angular scales was interpreted as a clear hint to a non-trivial topology of our Universe. The first findings of NASA's explorer mission ``Wilkinson Microwave Anisotropy Probe'' (WMAP) \\cite{Hinshaw_et_al_2003} have tremendously increased our knowledge of the temperature fluctuations of the CMB, since WMAP has measured the anisotropy of the CMB radiation over the full sky with high accuracy. The WMAP data confirm not only COBE's measurement of the low quadrupole moment $C_2$, but display in the temperature (auto-) correlation function $C(\\vartheta)$ $\\,(\\hat n \\cdot \\hat n' = \\cos \\vartheta)$ \\begin{equation} \\label{Eq:C_theta} C(\\vartheta) \\; \\simeq \\; \\frac 1{4\\pi} \\, \\sum_{l=2}^\\infty \\, (2l+1) \\, C_l \\, P_l(\\cos\\vartheta) \\end{equation} very weak correlations at wide angles, $70^\\circ \\lesssim \\vartheta \\lesssim 150^\\circ$, see the dashed curve in figure \\ref{Fig:C_theta_concordance}. (Note that the monopole and dipole are not included in the sum (\\ref{Eq:C_theta}).) At the largest angles, $\\vartheta \\gtrsim 160^\\circ$, the WMAP-data display even a ``correlation hole'', i.\\,e.\\ negative values of $C(\\vartheta)$. In figure \\ref{Fig:C_theta_concordance} we also show as a dotted curve the theoretical prediction according to the concordance model using the best-fit values for the cosmological parameters as obtained by WMAP \\cite{Hinshaw_et_al_2003}. It is seen that the concordance model, i.\\,e.\\ the best-fit $\\Lambda$CDM model, does not reproduce the experimentally observed suppression at $\\vartheta \\gtrsim 60^\\circ$ and the observed correlation hole. Since the correlation function $C(\\vartheta)$ emphasizes large angular scales and thus the low $l$-range, it is an ideal indicator function to search for a fingerprint of a possible non-trivial topology of the Universe. \\begin{figure}[htb] \\begin{center} \\hspace*{-12pt}\\begin{minipage}{9cm} \\includegraphics[width=9cm]{psplots/C_theta_concordance.ps} \\put(-1,4){$\\vartheta$} \\put(-274,150){$C(\\vartheta)$} \\put(-276,130){$[\\mu \\hbox{K}^2]$} \\end{minipage} \\vspace*{-10pt} \\end{center} \\caption{\\label{Fig:C_theta_concordance} The correlation function $C(\\vartheta)$ from the WMAP-data (dashed curve) in comparison with the concordance model (dotted curve). } \\end{figure} The evidence for a low quadrupole and missing power on large scales is, however, currently heavily discussed. The low value of the quadrupole obtained by the WMAP team \\cite{Hinshaw_et_al_2003} has been criticized by the way how the foregrounds \\cite{Bennett_et_al_2003b} are taken into account which are mainly caused by free-free, synchrotron and dust emission. Alternative reconstructions of the true cosmological signal from the five frequency bands measured by WMAP are discussed in \\cite{Tegmark_deOliveira_Costa_Hamilton_2003,Eriksen_Banday_Gorski_Lilje_2004} which lead to a higher quadrupole moment. The observation that the plane of the quadrupole and two of the three planes of the octopoles are aligned towards the ecliptic might be seen as a hint of an unknown source or sink of CMB radiation in the outer solar system or as an unrecognized systematic \\cite{Schwarz_Starkman_Huterer_Copi_2004}. The influence of the approximation of the likelihood function used for the angular power spectra analysis is investigated in \\cite{Slosar_Seljak_Makarov_2004}, where for the lowest multipoles an exact likelihood function estimation is presented. The statistical significance of the low value of the quadrupole moment is also discussed in \\cite{Efstathiou_2003} who found the WMAP results to be consistent with the concordance $\\Lambda$CDM model. The statistic is limited by the one-sky realization we are able to measure, also called cosmic variance. To circumvent this problem it is suggested to utilize the fact that the CMB polarization is sourced by the local temperature quadrupole. On the one hand, the polarization signal contains information about the quadrupole moment at the reionization epoch. This could lead to a probability for the observed quadrupole moment of the order of $10^{-4}$ \\cite{Skordis_Silk_2004} compared with the concordance model. On the other hand, future measurements of the linear polarization of the CMB towards clusters of galaxies can betray the local quadrupole moment at the location of the cluster \\cite{Kamionkowski_Loeb_1997,Sazonov_Sunyaev_1999,Seto_Sasaki_2000} thus circumventing the cosmic variance limit. Furthermore, we would like to mention that ``the question whether the geometry of the three-dimensional space of astronomy might be non-Euclidean'' \\cite{Chandrasekhar_1986} has already been posed by Schwarzschild \\cite{Schwarzschild_1992,Schwarzschild_1900,Schwarzschild_1998} in 1900, fifteen years before the founding of general relativity! He stated the problem as follows (see p.\\,32 in \\cite{Schwarzschild_1992}): ``As must be known to you, during this century [meaning the 19th century] one has developed non-Euclidean geometry (besides Euclidean geometry), the chief examples of which are the so-called spherical and pseudo-spherical spaces. We can wonder how the world would appear in a spherical or pseudo-spherical geometry with possibly a finite radius of curvature \\dots One would then find oneself, if one will, in a geometrical fairyland; and one does not know whether the beauty of this fairyland may in fact be realized in nature.'' Schwarzschild ``actually estimated limits to the radius of curvature of the three-dimensional space with the astronomical data available at his time and concluded that if the space is hyperbolic its radius of curvature cannot be less than 64 light years and that if the space is spherical its radius of curvature must at least be 1600 light years'' \\cite{Chandrasekhar_1986}. In an Addendum to his article ``On the permissible scale of curvature of space'' \\cite{Schwarzschild_1900} he mentioned also the possibility of spaces with non-trivial topology by referring to Clifford-Klein space forms. And he emphasized that such spaces do not necessarily lead to infinite universes as commonly assumed, even in the case of Euclidean or hyperbolic geometry. Schwarzschild concluded that the only condition imposed by astronomical data is that the volume of the Universe must be larger than the visible system of stars. It is the purpose of this paper to investigate two models of the Universe whose global topology is not the universal covering space of their spatial geometry. The spatial geometry of the models is the pseudo-spherical or hyperbolic space, i.\\,e.\\ it has negative curvature. The corresponding hyperbolic universes are non-compact and possess an infinitely long horn. The first model was introduced in 1976 by Sokolov and Starobinskii \\cite{Sokolov_Starobinskii_1976} and is described by a fundamental cell ${\\cal F}$ having infinite volume. The second model, called the Picard model \\cite{Picard_1884}, has also an infinitely long horn, but nevertheless possesses a finite volume. The Picard model has been investigated in detail in the context of quantum chaos \\cite{Matthies_1995,Steil_1999,Then_2003}. The Sokolov-Starobinskii model has been studied already in \\cite{Sokolov_Starobinskii_1976} and \\cite{Levin_Barrow_Bunn_Silk_1997} (see also the review \\cite{Levin_2002}). In \\cite{Levin_Barrow_Bunn_Silk_1997} it has been claimed that the periodic horn topology produces a flat spot in the temperature map of the CMB even when multiple images of astronomical sources are unobservable. The flat spot, i.\\,e.\\ a large flat region in the CMB sky map corresponding to a negligibly small metric perturbation, is argued in \\cite{Levin_Barrow_Bunn_Silk_1997} to result from the exponential decay of the eigenmodes in the direction of the horn. We shall show in the following that the flat spot found in \\cite{Levin_Barrow_Bunn_Silk_1997} is a consequence of a too low wavenumber cut-off $k_c$. For the cut-off chosen in \\cite{Levin_Barrow_Bunn_Silk_1997}, the eigenmodes cannot produce indeed a perturbation in the horn at the position which corresponds to the distance of the surface of last scattering (SLS). As a consequence, the sky maps computed in \\cite{Levin_Barrow_Bunn_Silk_1997} do not show temperature fluctuations in the horn. However, if the cut-off is sufficiently increased, we shall demonstrate that there is no suppression in the horn and therefore no flat spot. The universes described by the Sokolov-Starobinskii and Picard model, respectively, possess negative spatial curvature in contrast to the concordance model corresponding to a spatially flat universe. Therefore the question arises whether a negatively curved universe is realized in nature. In \\cite{Aurich_Steiner_2002b,Aurich_Steiner_2003} we have analysed the CMB data and the magnitude redshift relation of supernovae type Ia in the framework of quintessence models and have shown that these data are consistent with a nearly flat hyperbolic geometry of our Universe if the optical depth $\\tau$ to the SLS is not too big. The restriction comes from the large amplitude of the fluctuations at large scales in the CMB. However, it was shown in \\cite{Conversi_Melchiorri_Mersini_Silk_2004} that by replacing the quintessence component having a rest frame sound velocity of $c_s = c$ by a generalized dark matter component \\cite{Hu_1998} with a vanishing rest frame sound velocity $c_s = 0$, the amplitude in the CMB anisotropy is much smaller at large scales. Thus with such a generalized dark matter component universes with negative curvature are permissible even for larger optical depths $\\tau$. Recently, an ellipticity analysis of the CMB maps has been reported \\cite{Gurzadyan_et_al_2003a,Gurzadyan_et_al_2003b,Gurzadyan_et_al_2004} which gives further support to a hyperbolic spatial geometry of the Universe. In these analyses hot and cold anisotropy spots in the CMB maps have been studied in terms of shape for various temperature thresholds. Analysing with the same algorithm the COBE-DMR, BOOMERanG 150 GHz and WMAP maps, an ellipticity of the anisotropy spots has been found of the same average value (around 2) for these experiments. The WMAP data confirm the effect for scales both smaller and larger than the horizon at the SLS. This suggests that the effect is not due to physical effects at the SLS, and can arise after, while the photons are moving freely in the Universe. Finally, we would like to mention that recently a model was presented \\cite{Luminet_Weeks_Riazuelo_Lehoucq_Uzan_2003} possessing positive curvature and a finite volume with the shape of the Poincar\\'e dodecahedral space. The authors of ref.\\ \\cite{Luminet_Weeks_Riazuelo_Lehoucq_Uzan_2003} calculated the CMB multipoles for $l=2,3$ and 4, set the overall normalization factor to match the WMAP data at $l=4$ and examined the prediction for $l=2$ and 3. They found a weak suppression of the power at $l=3$ and strong suppression at $l=2$ in agreement with the WMAP observations. However, in ref.\\ \\cite{Luminet_Weeks_Riazuelo_Lehoucq_Uzan_2003} only the modes up to $k_c=30$ have been used, and it thus remains the question about how this low cut-off affects the integrated Sachs-Wolfe (ISW) contribution. Our experience shows that increasing the cut-off usually enhances the ISW contribution. ", "conclusions": "A large part of this paper was devoted to the question of the existence of flat spots in the CMB sky maps for universes with a horned topology. By the example of two such universes, the Sokolov-Starobinskii model and the Picard model, we showed that the infinitely long horn does not lead to flat spots, i.\\,e.\\ to a suppression of the CMB fluctuations in the horn, if the wavenumber cut-off $k_c$ is chosen sufficiently large. The flat spots reported earlier \\cite{Levin_Barrow_Bunn_Silk_1997,Levin_2002} are thus seen as the result of taking not enough modes into account in the expansion of the metric perturbation. We conclude that the CMB sky maps do not reveal a signature of the horned topology in form of flat spots. Another main point of this paper was to show that a universe with a horned topology, but with a finite volume, such as the Picard model, can explain the loss of power at large angular scales in the CMB anisotropy, as observed by COBE and WMAP. However, if the volume is infinite, as in the Sokolov-Starobinskii model, we have demonstrated that the low quadrupole moment is not reproduced. There is, however, some controversy of how serious one has to take this low value. The small values of the first few multipole moments $C_l$ have, on the one hand, not been taken so serious to require some new physics, but rather have been considered as a manifestation of cosmic variance or an unsufficient consideration of Galactic emission \\cite{Efstathiou_2003} or caused by the local supercluster \\cite{Abramo_Sodre_2003}. From this point of view the Sokolov-Starobinskii model is a viable model. On the other hand, the suppression has been taken as a hint to new physics such as new information on the inflation potentials \\cite{Cline_Crotty_Lesgourgues_2003,% Contaldi_Peloso_Kofman_Linde_2003,Feng_Zhang_2003}. However, these potentials have to be fine-tuned such that the arising power spectra are suppressed around the present day cosmological horizon \\cite{Contaldi_Peloso_Kofman_Linde_2003}. Such a fine-tuning is not necessary in the case of a non-trivial topology in a universe with negative spatial curvature due to the Mostow rigidity theorem \\cite{Mostow_1973,Prasad_1973}. Thus if the curvature scale is fixed by the densities $\\Omega_x$, all side-lengths of the fundamental cell ${\\cal F}$ are determined, and in turn the comoving wavenumber $k$ at which the spectrum is suppressed, not because of the initial power spectrum but simply because of the absence of modes below the lowest wavenumber $k_1$. In models with a non-trivial topology, certain points at the surface of last scattering can be identical due to the periodicity condition. These matching points are located on pairs of circles with the same radius. Along two such circles the temperature fluctuations $\\delta T$ produced by the naive Sachs-Wolfe effect are the same, which is called the circles-in-the-sky signature \\cite{Cornish_Spergel_Starkman_1998b}. In \\cite{Cornish_Spergel_Starkman_Komatsu_2003} a search for such circles in the WMAP data was carried out for nearly back-to-back circles, i.\\,e.\\ for circles whose centers have a distance greater than 170$^\\circ$ and whose radii are greater than 25$^\\circ$ on the sky. They found no signature and rule out all topologies having such circles. However, it should be kept in mind that only the naive Sachs-Wolfe contribution leads to identical temperature fluctuations. The integrated Sachs-Wolfe contribution arises on the photon path to the observer, which is not identified for the observer and the ``copy'' of the observer. The same is valid for the Doppler contribution, since the observer and its copy see another projection of the velocity, in general. Furthermore, there are a lot of other secondary contributions to the temperature fluctuations. Nevertheless, it is claimed in \\cite{Cornish_Spergel_Starkman_Komatsu_2003} that the naive Sachs-Wolfe contribution is strong enough for the identification of circles-in-the-sky. The Picard model is not ruled out by this work, since there are no nearly back-to-back circles. For the model with $\\Omega_{\\hbox{\\scriptsize mat}} = 0.3$ and $\\Omega_\\Lambda = 0.65$ with $x_3=1.6$, we obtain 40 pairs, where the largest distance of the centers is at 145$^\\circ$ which is not covered by the study in \\cite{Cornish_Spergel_Starkman_Komatsu_2003}. The model with $\\Omega_{\\hbox{\\scriptsize mat}} = 0.35$ and $\\Omega_\\Lambda = 0.6$ has only 32 circle pairs. The number of paired circles increases if the observer is posited higher up in the horn, i.\\,e.\\ at larger values of $x_3$. For the extreme position at $x_3=5$ one obtains 275 circle pairs with distances up to 168.6$^\\circ$ for $\\Omega_{\\hbox{\\scriptsize mat}} = 0.3$ and $\\Omega_\\Lambda = 0.65$. These circles could have been detected in \\cite{Cornish_Spergel_Starkman_Komatsu_2003}. However, for more generic observers, which are not sitting extremely high in the horn, the separation of the circles is too small such that the model cannot yet be ruled out. In conclusion, we would like to emphasize again that the Picard model studied in this paper is in nice agreement with the observed suppression of power on large scales in the angular power spectrum of the CMB (see figures \\ref{Fig:Picard_gen_ran_m30_q00_l65_h65_C_theta}, \\ref{Fig:Picard_gen_ran_m35_q00_l60_h65_C_theta}, \\ref{Fig:Picard_gen_Spitze_ran_m30_q00_l65_h65_C_Theta_cusp} and \\ref{Fig:Picard_gen_ran_m30_q00_l65_h65_Merge_C_Theta}). This is in contrast to the concordance model which does not reproduce the experimentally observed suppression at $\\vartheta \\gtrsim 60^\\circ$ and the observed correlation hole at $\\vartheta \\gtrsim 160^\\circ$. If future observations will confirm the WMAP data but with smaller errors, this can be interpreted as a clear hint to a non-trivial topology of our Universe having negative spatial curvature and a finite volume." }, "0403/astro-ph0403242_arXiv.txt": { "abstract": "We present the first determination of the 15$\\mu$m luminosity function of galaxies from the European Large Area ISO survey (ELAIS) southern fields. We have adopted a new criterion to separate the quiescent, non-evolving and the starburst, evolving populations based on the ratio of mid-infrared to optical luminosities. Strong evolution is suggested by our data for the starburst galaxy population, while normal spiral galaxies are consistent with no evolution. The starburst population must evolve both in luminosity and in density with rates of the order $L(z) \\propto (1+z)^{3.5}$ and $\\rho(z) \\propto (1+z)^{3.8}$ up to $z \\sim 1$. The evolutionary parameters of our model have been tested by comparing the model predictions with other observables, like source counts at all flux density levels (from 0.1 to 300 mJy) and redshift distributions and luminosity functions at high-$z$ (0.7 $< z <$ 1.0 from HDF-N data). The agreement between our model predictions and the observed data is remarkably good. We use our data to estimate the star-formation density of the Universe up to z=0.4 and we use the luminosity function model to predict the trend of the star-formation history up to $z=1$. ", "introduction": "\\label{intro_sec} The extragalactic background light shows that the emission from galaxies at infrared and sub-millimeter wavelengths is an energetically significant component of the Universe. This emission originates from star-formation activity and active galactic nuclei. The precise contribution from each type of activity is still debated. It is thus important to our understanding of galaxy and AGN formation to study those populations that emit a substantial amount of light at infrared (IR) wavelengths in their rest frame. In particular, data from deep {\\it Infrared Space Observatory} (ISO) surveys at 15 $\\mu$m (i.e. \\citealp{1999A&A...351L..37E}; \\citealp{1999ApJ...517..148F}; \\citealp{2001MNRAS.325.1173L}; \\citealp{2003A&A...407..791M}) seem to require strong evolution for galaxies emitting in the infrared wavebands. This result, supported also by the detection of a substantial cosmic Infrared Background in the 140 $\\mu$m - 1 mm range \\citep{1996A&A...308L...5P,1998ApJ...508...25H,1999A&A...344..322L}, has stimulated the development of several evolutionary models for IR galaxies (i.e. \\citealp{2001ApJ...549..745R}, \\citealp{2001A&A...378....1F}, \\citealp{2001ApJ...556..562C}, \\citealp{2003ApJ...587...90X}). All these models fit with different degrees of success the IR/sub-millimeter source counts and the cosmic Infrared Background within the present uncertainty limits, but suffer of parameter degeneracy and none of them is based on a local luminosity function (LLF) obtained from 15-$\\mu$m data, being all extrapolated from different IR wavelengths (12, 25, 60 $\\mu$m). So far, complete spectroscopic samples of 15-$\\mu$m sources have been obtained only in small fields (i.e. HDF-N: \\citealp{1999usis.conf.1023A}; HDF-S: \\citealp{2002MNRAS.332..549M}, \\citealp{2003A&A...403..501F}), too small and too deep to allow a detailed study of the local luminosity function. The ELAIS survey is the largest open time project conducted by ISO \\citep{2000MNRAS.316..749O}, mapping an area of $\\approx$ 12 deg$^2$ at 15 $\\mu$m and 90 $\\mu$m. The final, band-merged ELAIS catalogue has recently been completed \\citep{MRR03}. The spectroscopic data is most complete in the southern fields and in this paper we present an analysis of the 15-$\\mu$m luminosity function derived from these data. This is the first determination of the 15-$\\mu$m luminosity function and its evolution, constrained by all the available observables in this band (source counts from IRAS to the deepest ISOCAM flux densities; redshift distributions at low and high-$z$ using both data from ELAIS and data from the deeper HDF-N survey). The model fitting the LF is then used to estimate the star-formation history of the Universe. This paper is structured as follows. In Section \\ref{sample_sec} we present our data sample. In Section \\ref{kcorr_sec} we discuss the adopted IR and optical K-corrections. In Section \\ref{fdl_sec} we show the method used to compute the 15-$\\mu$m LF and present the results. In Section \\ref{llf_section} we compare our LLF determination with previous ones. In Section \\ref{evolution_sec} we discuss the evolution rates derived from our data, and compare the model predictions with the observable constraints and with other literature models. In Section \\ref{concl_sec} we present our conclusions. Throughout this paper we will assume $H_0 = 75$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_m = 0.3$ and $\\Omega_{\\Lambda} = 0.7$, unless explicitly stated. ", "conclusions": "\\label{concl_sec} We have presented the first direct determination of the 15-$\\mu$m luminosity function and its cosmic evolution for galaxies. As previously found by other authors, three populations of sources give rise to the 15-$\\mu$m emission: the normal, the starburst and the AGN populations, characterized by different cosmic evolution. In this work we have analysed the galaxy component only (quiescent plus actively starforming). The contribution of the AGN component is discussed in a separate paper (\\citealp{2002MNRAS.332L..11M}; Matute et al. in preparation). The analysis is based on data from the ELAIS southern fields survey. The sample is composed by $\\sim$150 galaxies in the redshift interval 0.0$<$z$\\le$0.4 and covers a large flux density range intermediate between the IRAS and the deep ISOCAM surveys (0.5$\\le{S}\\le{50}$ mJy). Differently from other authors, we have adopted in this work the $L_{15{\\mu}m}/L_{\\rm R}$ ratio as a criterion to separate the quiescent, non-evolving and the starburst, evolving populations. This criterion, suggested by the existing correlation between $L_{15{\\mu}m}/L_{\\rm R}$ and the amount of activity in galaxies, is a posteriori supported by the results of the V/V$_{max}$ analysis on the two populations defined on this basis. The main results of our analysis are: \\begin{enumerate} \\item{In the ML analysis we have simultaneously fitted both the evolution rates and the shape parameter of the local LF for both the spiral and the starburst populations. We have assumed that the spiral population does not evolve, while we have let to evolve the starburst population both in luminosity and in density. Since the two populations sample different luminosity ranges, we have obtained an accurate determination for the faint end of the LLF for the quiescent component, while the knee and the $\\sigma$ parameters of the LLF are better constrained for the starburst one. The evolution found for the active population is $\\sim(1+z)^{3.5}$ in luminosity and $\\sim(1+z)^{3.8}$ in density, up to $z_{break}\\sim1$.} \\item{Our total 15-$\\mu$m LLF is in agreement with previous determinations derived from the IRAS data. On the contrary, the LLFs for different populations derived by different authors have not the same level of consistency. While in our subdivision the quiescent population is expected to dominate locally over all the luminosity range, in other models (i.e. \\citealp{2003ApJ...587...90X}) the starburst population dominates locally at high luminosities, leading to a large discrepancy in the model predictions.} \\item{To test the evolution parameters with higher degree of confidence, we have compared our model predictions with all the observables existing in literature, over all the $z$ and flux ranges (source counts, luminosity functions, $z$-distributions). Our best-fitting model well reproduces all the observables. In the critical interval 1${\\lsimeq}S{\\lsimeq}$3 mJy, where the source counts from different surveys show the larger discrepancies, our model is intermediate between the data from ELAIS-S1 \\citep{2002MNRAS.335..831G} and the data from the deep surveys \\citep{1999A&A...351L..37E}. On the other hand, in the flux range 3${\\lsimeq}S{\\lsimeq}$10 mJy, our model is well consistent with existing data, differently from the \\cite{2003ApJ...587...90X} model, whose predicted differential sources counts are at least a factor of 3 higher than the data.} \\item{Using the evolutionary model found for the 15-$\\mu$m galaxies and the data points from the $1/V_{max}$ LF analysis, we have estimated the star-formation rate density. The redshift range sampled by our data ($0.0 < z \\le0.4$) is of particular interest, since it provides an estimate of the star-formation at relatively low redshift, but not so local to be affected by clustering and local dishomogeneities. We find $\\dot \\rho$=(0.025$\\pm$0.007) hM$_\\odot$yr$^{-1}$Mpc$^{-3}$ and $\\dot \\rho$=(0.043$\\pm$0.020) hM$_\\odot$yr$^{-1}$Mpc$^{-3}$ at the two mean redshifts $z=0.12$ and $z=0.27$, respectively. At $z \\lsimeq 0.4$ our model predictions are well consistent with other estimates derived from UV, optical and Mid-IR data. At higher redshift our model predictions are significantly higher than the UV extinction corrected data and lower by about a factor of two than the estimates derived from radio data by \\cite{2000ApJ...544..641H}.} \\end{enumerate}" }, "0403/gr-qc0403080_arXiv.txt": { "abstract": "In order to detect the rare astrophysical events that generate gravitational wave (GW) radiation, sufficient stability is required for GW antennas to allow long-term observation. In practice, seismic excitation is one of the most common disturbances effecting stable operation of suspended-mirror laser interferometers. A straightforward means to allow more stable operation is therefore to locate the antenna, the ``observatory'', at a ``quiet'' site. A laser interferometer gravitational wave antenna with a baseline length of 20\\,m (LISM) was developed at a site 1000\\,m underground, near Kamioka, Japan. This project was a unique demonstration of a prototype laser interferometer for gravitational wave observation located underground. The extremely stable environment is the prime motivation for going underground. In this paper, the demonstrated ultra-stable operation of the interferometer and a well-maintained antenna sensitivity are reported. ", "introduction": "First-generation ground-based gravitational wave antennas (LIGO-I~\\cite{ligo1, ligo2}, VIRGO~\\cite{virgo1, virgo2}, GEO\\,600~\\cite{geo1, geo2, geo3}, TAMA\\,300~\\cite{tama1, tama2, tama3}) are expected to come on-line early in this decade as a global network searching for astrophysical gravitational wave radiation. At present, some of the detectors are already operating intermittently, hoping to observe the spacetime strain of the universe. The aim of these international projects is to directly detect gravitational radiation, faint ripples in the spacetime fabric. There are several kinds of expected astrophysical sources, including chirping gravitational waves from inspiraling compact star binaries, burst signals from supernovae explosions, and the stochastic background radiation. The expected event rate of these sources is, however, quite low even if the uncertainty of the population estimate~\\cite{eventrate1, eventrate2, eventrate3} is taken into account, so, to avoid missing these rare and faint signals, stable operation of the detector, keeping the duty cycle and the detector sensitivity high and also keeping the data quality high, are indispensable requirements for a gravitational wave observatory. In general, the technologies used in a laser interferometer are based on an ultra-high precision measurement pursuing extremely high sensitivity, so the instruments are very sensitive to almost all kinds of noise, disturbances and drifts. The noise source that most commonly disturbs stable operation of suspended-mirror laser interferometers is seismic excitation. The most promising solution for this problem is to avoid the source of these disturbances by selecting a quiet environment for a detector site. The goal of this project (LISM, Laser Interferometer gravitational-wave Small observatory in a Mine) is to demonstrate stable operation of the laser interferometer and to obtain high quality data for searching for gravitational waves at a well-suited observatory site. The 20-m baseline laser interferometer was originally developed for various prototyping experiments~\\cite{20m1,20m2,20m3,20m4} at the campus of The National Astronomical Observatory of Japan, in Mitaka, a suburb of Tokyo from 1991 to 1998. In 1999, it was moved to the Kamioka mine in order to perform long-term, stable observations as LISM. In this paper, the merit of going underground and the demonstrated stable operation of the antenna are reported. The results of data analysis and a GW search will appear as a separate article~\\cite{coincidence}. ", "conclusions": "The prototype laser interferometer gravitational wave antenna LISM with an arm length of 20\\,m was moved into a deep underground laboratory, a first for a gravitational wave observatories, and a long-term observational run was performed. The goal of this project was to confirm that the laser interferometer can operate stably and provide high quality data, making the most of the stable environment in the underground laboratory. The interferometer sensitivity achieved was $1\\x10^{-18}\\rm\\,m/\\sqrt{Hz}$ around 800\\,Hz in displacement, which corresponds to $5\\x10^{-20}\\rm/\\sqrt{Hz}$ in strain. At higher frequencies, above that floor sensitivity, LISM was quantum noise limited by shot noise. With this sensitivity spectrum, LISM can detect gravitational wave events emitted from coalescence of 1.4--1.4\\,$M_{\\odot}$ equal-mass binary systems a few\\,kpc away by matched filtering analysis with a SNR of 10. The operational duty cycle of the interferometer exceeded 99.8\\% owing to the stable, low-disturbance environment and the self-recovering automation systems. If we adopt the expected SNR of the matched filtering analysis as an index for stability of the interferometer sensitivity, this SNR duty cycle was about 90\\%. This means that the interferometer sensitivity was kept within a 3\\,dB window for 90\\% of the observational period. According to these results, we conclude that the laser interferometer gravitational wave antenna was operated stably enough for a long term observational run, and the underground environment is suitable as a gravitational wave antenna site." }, "0403/astro-ph0403524_arXiv.txt": { "abstract": "We have mapped the ultracompact \\hii\\ region, G5.89-0.39, and its molecular surroundings with the Submillimeter Array at $2\\arcseconds.8 \\times 1\\arcseconds.8$ angular resolution in 1.3 mm continuum, \\siofive, and eight other molecular lines. We have resolved for the first time the highly energetic molecular outflow in this region. At this resolution, the outflow is definitely bipolar and appears to originate in a 1.3 mm continuum source. The continuum source peaks in the center of the \\hii\\ region. The axis of the outflow lines up with a recently discovered O5V star. ", "introduction": "\\label{section:int} G5.89-0.39 (G5.89) is a shell-like ultracompact (UC) \\hii\\ region \\citep{woo89b} which was recently found to contain an O5V star via near-IR imaging \\citep{fel03}. \\citet{aco98} have used proper motion measurements of the expansion of G5.89 to determine its distance, $\\sim 2$ kpc, its size, 0.01 pc, and its dynamical age, 600 years. Associated with G5.89 is a molecular outflow \\citep{har88} with a mass of 77 \\msun\\ and an energy of $5 \\times 10^{47}$ ergs \\citep[hereafter, AWC]{aco97}. While many \\uchii\\ regions have been observationally associated with molecular outflows \\citep{sne90,she96a}, the question of whether an \\uchii\\ region can itself be or contain the source of a bipolar outflow remains in dispute. Bipolar outflows are generally understood to be a signpost of ongoing accretion, while the presence of an \\uchii\\ region has been understood by some as a sign that accretion has ceased \\citep{gar99}. Recent observational and theoretical results have shown that the presence of an \\uchii\\ region does not necessarily shut off accretion \\citep{ket02a,ket02b}. But since massive stars and their \\uchii\\ regions tend to form in very crowded fields, high resolution imaging is necessary to associate the origin of any particular bipolar molecular outflow with an \\uchii\\ region, and not a distinct, nearby protostar. This determination is possible with and well suited to the capabilities of the new Submillimeter Array\\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. For more on the SMA and its specifications see \\citet{ho03}.} (SMA) on Mauna Kea. We have used the SMA to map the region of G5.89-0.39 in the \\siofive\\ line, known to trace outflows and this outflow in particular \\citep[AWC]{cod99}. Because of the spectral capabilities of the SMA, we were simultaneously able to map the region in nine spectral lines and broadband continuum. The aim of the project was to improve on the previous $18\\arcseconds$ resolution map in \\siofi\\ of AWC, and also to detect in dust and molecular lines any other massive protostars in the vicinity which might be sources of the outflow. ", "conclusions": "\\label{section:discussion} We believe the origin of the outflow is the 1.3 mm continuum source for two reasons. First, the slightly extended 1.3 mm continuum peak lies on the projected outflow axis, halfway between the peaks of the two outflow lobes. Second, the 1.3 mm continuum source seems to show a physical connection to both outflow lobes in the form of the extensions of the continuum along the outflow axis. Apparently the dust is being swept up in both lobes of the outflow (as seen in L1157 by \\citet{gue03}), and thus it seems that the outflow is emanating from the 1.3 mm continuum source. The question of whether the outflow actually originates {\\it inside} the \\uchii\\ region is somewhat thornier. One possibility is that the 1.3 mm continuum source lines up with the \\uchii\\ region by chance, and is a distinct high-mass protostar with its own dusty envelope, not physically connected to the \\uchii\\ region, but part of the same young cluster. Alternatively, the dust and free-free continuua could be from a single source, one dusty, hot molecular core surrounding an embedded \\uchii\\ region. In this case, it seems the outflow would have to come from inside the \\uchii\\ region in order for both sides of the outflow to be seen in the dust continuum. We cannot distinguish between these two cases, and neither seems unlikely. However, we can say that the known O5 star is a bad candidate to be the source of the outflow since it is not equidistant from the outflow lobes. We cannot eliminate any model chronologies by comparing timescales. The outflow has a dynamical age of 1600 years, compared to just 600 for the \\uchii\\ region. This might lead one to believe that the outflow is truly older than the \\uchii\\ region. However, any ``quenching'' of the \\uchii\\ region by infalling material would tend to make the dynamical age of the \\hii\\ region an underestimate of its true age, perhaps by a great deal \\citep{ket02b}. The gas in the outflow was probably accelerated before the \\uchii\\ region began to expand, but not necessarily before the \\uchii\\ region was created. The central unknown about the outflow is its inclination. The fact that the two lobes hardly overlap at all on the plane of the sky and the fact that the \\siofi\\ emission is extended perpendicular to the outflow axis might lead one to believe that the outflow has a wide opening angle, and lies primarily in the plane of the sky. But these two facts could just as easily be interpreted as a highly collimated outflow, like those seen around low-mass proto-stars \\citep{bac91}, whose axis is largely along the line of sight. Including only the line-of-sight velocity, the outflow is already quite energetic, so adding a large component in the plane of the sky would make this a remarkable outflow, indeed. Some studies of regions of massive star formation have found poor collimation of outflows \\citep{ric00,rid01}. But \\citet{beu02a} found that 15 outflows with apparently low collimation were consistent with the very high collimations seen in low mass cases, with the discrepancy due to low resolution. Higher resolution interferometer studies have confirmed the presence of highly collimated outflows in some regions of high-mass star formation \\citep{beu02b}. The velocity structure of the \\siofi\\ is intriguing and may be a clue as to the outflow orientation. In both lobes, the spectra peak at low relative velocity and tail off towards much higher relative velocity. In the case of outflows from low-mass protostars, the SiO emission is concentrated immediately behind the shock in the highest velocity molecular gas \\citep{van98}. So perhaps the fact that the low relative velocity peaks in the \\siofi\\ emission, only $\\pm 5$ \\kms, is an indication that the outflow is mainly in the plane of the sky. This outflow, however, is far more energetic than an outflow from a low-mass protostar. There could be highly excited SiO emission which follows behind the shock more closely than the \\siofi\\ emission. We plan to observe \\sioeight, and lines of several CO isotopes with the SMA allowing calculations of excitation and abundance in the outflow to see if higher excitation gas occurs at higher line of sight velocities, as we might expect for an outflow largely along the line of sight. These observations will also have higher angular resolution which should clarify the outflow opening angle. On an equally hypothetical note, consider the fact that while some dust emission appears to come from the outflow, the free-free emission shows no preference for the direction of the outflow, and is more extended perpendicular to the outflow. Recent results have shown that massive protostars ($\\sim 10$\\msun) can have disk-like structures akin to the disks seen in low-mass protostars \\citep{zha98b,zha02}. If such a disk were to photoevaporate, as was modeled by \\citet{hol94}, that would naturally produce free-free emission which is extended perpendicular to the outflow, and possibly ring-like. However, if the outflow does not originate in the \\uchii\\ region, the more traditional limb-brightened-shell interpretation of the free-free emission may make more sense. Future work to investigate the geometry and the velocity structure of the ionized gas could be carried out with the SMA using low quantum number radio recombination lines (RRLs) of hydrogen, including \\htwonea, \\htwsixa, \\htha, and \\hthonea. These lines suffer less from pressure broadening than higher quantum number RRLs, and could be used to probe the velocity structure of the densest ionized gas in the \\uchii\\ region." }, "0403/astro-ph0403238_arXiv.txt": { "abstract": "{ The accretion column in magnetic Cataclysmic Variables may have a not negligible Thomson optical depth. A fraction of the thermal radiation from the post--shock region may therefore be scattered -- and then polarized -- before escaping the column. Moreover, part of the thermal radiation is reflected -- and again polarized -- by the White Dwarf surface. We show that X--ray polarimetry can provide valuable, and probably unique informations on the geometry and physical parameters of the accretion column by calculating, by means of Monte Carlo simulations, the expected polarization properties of magnetic CVs as a function of the geometrical parameters (assuming a cylindrical geometry) and the Thomson optical depth of the column. We find that degrees of polarization as high as about 4\\% can be present, and apply our calculations to the archetypal magnetic CV, AM~Herculis. ", "introduction": "The magnetic field in some subclasses of Cataclysmic Variables (notably in Polars and in at least a fraction of intermediate Polars; see Warner 1995 for a complete overview on CVs) is strong enough to channel the accreting matter along the field lines. For a dipolar field, this means that the accretion occurs via accreting columns on one or both the magnetic poles; the disalignement between the spin and magnetic axes results in pulsed emission. Hard X--rays are then produced by optically thin thermal line and continuum emission in the so--called post--shock region, where temperatures can reach values as large as several tens of keV (e.g. Frank et al. 1992; Cropper et al. 1999). Thermal emission is expected to be unpolarized (see next section); however, the Thomson optical depth in the accretion column, while probably less than unity, may be not negligible. Indeed, Hellier et al. (1998) found significant broadening in the iron K$\\alpha$ line of several magnetic CVs, which they interpreted as due to Compton broadening. Thomson scattered radiation is polarized, provided that the geometry is not spherical. It is therefore to be expected that the hard X--ray emission in magnetic CVs is polarized, with the net polarization degree increasing with the Thomson optical depth of the accretion column. In this paper we calculate, by means of Monte Carlo simulations, the polarization properties of the accretion column in magnetic CVs. In Sec. 2 the numerical code is described, while the results (including an application to the archetypal Polar, AM~Herculis) are presented in Sec.~3. Results are then summarized, and observational perspectives discussed, in Sec.~4. \\begin{figure*}[t] \\hbox{ \\includegraphics*[width=8.0cm,height=8.0cm]{tau_0.1.ps} \\hspace{0.5cm} \\includegraphics*[width=8.0cm,height=8.0cm]{tau_0.2.ps} } \\hbox{ \\includegraphics*[width=8.0cm,height=8.0cm]{tau_0.5.ps} \\hspace{0.5cm} \\includegraphics*[width=8.0cm,height=8.0cm]{tau_1.ps} } \\caption{The degree of polarization, P, as a function of $\\mu=cos\\theta$ for four values of $H/r_C$. In each panel, five values of $\\tau_T$ are shown: 0.05, 0.1, 0.3, 0.5 and 1 (in increasing order of the absolute value of $P$). } \\label{polittico_1} \\end{figure*} \\begin{figure*}[t] \\hbox{ \\includegraphics*[width=8.0cm,height=8.0cm]{tau_2.ps} \\hspace{0.5cm} \\includegraphics*[width=8.0cm,height=8.0cm]{tau_3.ps} } \\hbox{ \\includegraphics*[width=8.0cm,height=8.0cm]{tau_5.ps} \\hspace{0.5cm} \\includegraphics*[width=8.0cm,height=8.0cm]{tau_10.ps} } \\caption{The same as in the previous figure, for other four values of $H/r_C$. } \\label{polittico_2} \\end{figure*} \\begin{figure*}[t] \\hbox{ \\includegraphics*[width=8.0cm,height=8.0cm]{tau_flux_0.1.ps} \\hspace{0.5cm} \\includegraphics*[width=8.0cm,height=8.0cm]{tau_flux_0.2.ps} } \\hbox{ \\includegraphics*[width=8.0cm,height=8.0cm]{tau_flux_0.5.ps} \\hspace{0.5cm} \\includegraphics*[width=8.0cm,height=8.0cm]{tau_flux_1.ps} } \\caption{The flux (in arbitrary units) as a function of $\\mu=cos\\theta$ for four values of $H/r_C$. In each panel, five values of $\\tau_T$ are shown: 0.05, 0.1, 0.3, 0.5 and 1 (in increasing order of anisotropy). } \\label{polittico_flux_1} \\end{figure*} \\begin{figure*}[t] \\hbox{ \\includegraphics*[width=8.0cm,height=8.0cm]{tau_flux_2.ps} \\hspace{0.5cm} \\includegraphics*[width=8.0cm,height=8.0cm]{tau_flux_3.ps} } \\hbox{ \\includegraphics*[width=8.0cm,height=8.0cm]{tau_flux_5.ps} \\hspace{0.5cm} \\includegraphics*[width=8.0cm,height=8.0cm]{tau_flux_10.ps} } \\caption{The same as in the previous figure, for four other values of $H/r_C$. } \\label{polittico_flux_2} \\end{figure*} \\medskip ", "conclusions": "We have calculated the polarization properties of the accretion column in magnetic CVs. Polarization arises because a fraction of the thermal radiation can be Thomson scattered before escaping the column. The polarization degree can be as high as $\\sim$4\\%, and depends on the angle between the magnetic field and the line--of--sight (which of course varies with the spin phase), which is often well known (e.g. Cropper 1988). It depends also on the Thomson optical depth, $\\tau_T$, and on the ratio between the radius and height of the accretion column. If one of these two parameters can be independently estimated (e.g. $\\tau_T$ from the iron line broadening, e.g. Hellier et al. 1998), than the other can in principle be deduced by polarization measurements. In particular, as the polarization degree is negative (positive) for $H/r_C$ less (greater) than one, while the polarization of the reflection component (whose importance increases with energy) is always negative, an increase (decrease) of the polarization degree with energy is expected in the former (latter) case. In this paper we have discussed the continuum emission. However, iron lines provide a significant fraction of the total X--ray flux. Recombination lines emitted in the accretion column may suffer not only Compton scattering but also resonant scattering, so the effective optical depth in the line is larger than in the continuum. Line photons are therefore likely to be more polarized than continuum photons. On the contrary, the fluorescent neutral line emitted by the White Dwarf surface is unpolarized, at least in the line core. Degrees of polarization of the order of one percent are within the detection capabilities of the new generation X--ray polarimeters based on the photoelectric effect (Costa et al. 2001) when coupled with large enough X--ray telescopes (Costa et al. 2003), at least for the brightest magnetic CVs in high state (when their flux can be as high as several millicrabs; for instance the 2--10 keV flux of AM~Herculis in high state is about 10$^{-10}$ erg cm$^{-2}$ s$^{-1}$, Matt et al. 2000). Moreover, the spectral resolution should be good enough to search for different polarization degrees in the emission iron lines. Bright magnetic CVs should therefore be added to the traditional lists of targets for future polarimetric missions." }, "0403/astro-ph0403712_arXiv.txt": { "abstract": "The Space Telescope Imaging Spectrograph has measured the absolute flux for Vega from 0.17--1.01~$\\mu$m on the \\emph{HST} White Dwarf flux scale. These data are saturated by up to a factor of 80 overexposure but retain linearity to a precision of 0.2\\%, because the charge bleeds along the columns and is recovered during readout of the CCD. The S/N per pixel exceeds 1000, and the resolution $R$ is about 500. A $V$~magnitude of 0.026$\\pm$0.008 is established for Vega; and the absolute flux level is $3.46\\times10^{-9}$~erg cm$^{-2}$ s$^{-1}$ at 5556~\\AA. In the regions of Balmer and Paschen lines, the STIS equivalent widths differ from the pioneering work of Hayes in 1985 but do agree with predictions of a Kurucz model atmosphere, so that the STIS flux distribution is preferred to that of Hayes. Over the full wavelength range, the model atmosphere calculation shows excellent agreement with the STIS flux distribution and is used to extrapolate predicted fluxes into the IR region. However, the IR fluxes are 2\\% low with respect to the standard Vega model of Cohen. \\emph{IUE} data provide the extension of the measured STIS flux distribution from 0.17 down to 0.12~$\\mu$m. The STIS relative flux calibration is based on model atmosphere calculations of pure hydrogen WDs, while the Hayes flux calibration is based on the physics of laboratory lamps and black body ovens. The agreement to 1\\% of these two independent methods for determining the relative stellar flux distributions suggests that both methods may be correct from 0.5--0.8~$\\mu$m and adds confidence to claims that the fluxes relative to 5500~\\AA\\ are determined to better than 4\\% by the pure hydrogen WD models from 0.12 to 3~$\\mu$m. ", "introduction": "The most commonly used flux calibration for the fundamental standard Vega ($\\alpha$~Lyr, HD~172167, HR~7001) is the compilation of Hayes (1985), while Megessier (1995) suggests an increase of 0.6\\% in the Hayes value of $3.44\\times10^{-9}$ erg cm$^{-2}$ s$^{-1}$ \\AA$^{-1}$ at 5556~\\AA. The Hayes flux for Vega is traceable to fundamental standard lamps pedigreed by NBS (now NIST), which is charged with maintaining fundamental standards of physical units. The Hayes flux compilation is based on direct comparisons of the star to calibrated lamps and black body ovens. Most ground based estimates of absolute stellar flux are traceable to the flux of Vega. With the advent of extensive space based observations in the far-UV, a need for standard flux candles arose at wavelengths below the atmospheric cutoff. A few rocket experiments established some UV standard stars with $\\sim$10\\% precision (e.g.\\ Strongylis \\& Bohlin 1979). To establish standards with better accuracy, a technique based on model atmospheres for simple, pure hydrogen WDs was suggested by D.~Finley and J.~Holberg for calibrating the \\emph{IUE} satellite spectrophotometry, (cf.\\ Fig.~8 in Bohlin, et~al.\\ 1990). The hot WD standard stars G191B2B, GD~153, and GD~71 at $V=12$--13~mag were established by computing LTE model atmosphere flux distributions for pure hydrogen atmospheres, which were then normalized to precisely measured $V$~magnitudes from A.~Landolt (see Bohlin 2000). EUV observations below the Lyman limit demonstrated that any effects of interstellar extinction in these three stars is less than 1\\% longward of the 912~\\AA\\ Lyman limit. These three WD standards are now based on NLTE models by Ivan Hubeny (Bohlin 2003) and are complemented by three solar analog stars for \\emph{HST} calibrations extending into the IR (Bohlin, Dickinson, \\& Calzetti 2001). The Vega observations with STIS are discussed in Section~2. Section~3 establishes the best estimate of the V band magnitude of Vega, while Section~4 presents the first direct comparison between the WD and the standard lamp methods for establishing absolute flux. A third method for defining the flux of Vega at all wavelengths is based on modeling of its stellar atmosphere, as discussed in Section~5. Model atmosphere codes are now sophisticated enough to warrant detailed comparisons between observation and theory. Because of the excellent agreement of a model (Kurucz 2003, Castelli and Kurucz 1994) with the STIS observations up to 1~$\\mu$m, the model should provide a good estimate of the flux beyond 1~$\\mu$m, as discussed in Section~6. ", "conclusions": "As a result of the STIS observations of Vega, a V mag of 0.026 is established for the improved bandpass function of Cohen et~al. \\ (2003b). The Kurucz \\ (2003) model of Vega agrees so well with the STIS observations that the model itself defines the standard star fluxes longward of 4200 \\AA. These improvements, combined with the more realistic bandpass function that is used to translate the V magnitudes of the WD standards into relative fluxes, have resulted in WD fluxes that are uniformly 0.5\\% fainter than found by Bohlin \\ (2000). The WD standard star model flux distributions must be updated in the CALSPEC archive, along with the new Vega spectrum. Because these files have not been updated since 00Apr4, the changes quantified by Bohlin (2003) as a function of wavelength for the switch from LTE to NLTE are also included in the new *mod*.fits files. The updates for the CALSPEC secondary flux standards are being prepared and will reflect the above changes plus a number of new STIS observations, better corrections for the changes with time and temperature, and improved CTE corrections for the CCD data. In addition, a suite of NICMOS grism calibration observations are being obtained to resolve the IR differences between the pure hydrogen WDs and the solar analogs and to extend the \\emph{HST} spectrophotometric coverage to 2.5~$\\mu$m. Unfortunately, Vega is too bright for NICMOS grism observations." }, "0403/astro-ph0403462_arXiv.txt": { "abstract": "We have performed \\HST\\ imaging of a sample of 23 high-redshift ($1.81.8$ AGN, all with nuclear luminosities near $M_B = -23$. In a companion paper \\citep{sanc04a} we study rest-frame colours and morphological properties of a sample of intermediate-redshift ($z \\la 1$) AGN. The paper is organised as follows. We first describe the sample selection and properties together with a summary of the observational data (Sect.~\\ref{sec:data}). We then comment on the decomposition of the nuclear and galaxy contribution, including a brief summary of the extensive simulations that we use to estimate measurement errors (Sect.~\\ref{sec:analysis}). In Sect.~\\ref{sec:results} we present the measured host galaxy magnitudes and describe our treatment of non-detections. We move on to discuss the results in Sect.~\\ref{sec:discussion}, followed by our conclusions in Sect.~\\ref{sec:conclusions}. We use $H_0=70$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_m=0.3$ and $\\Omega_\\Lambda = 0.7$ throughout this paper. All quoted magnitudes are zeropointed to the AB system with ZP$_\\mathrm{F606W}=26.493$ and ZP$_\\mathrm{F850LP}=24.843$. ", "conclusions": "\\label{sec:conclusions} We performed the hitherto largest study of host galaxies properties of a complete sample of high-redshift AGN. We detected the hosts and extract colour information in 9 of the 23 AGN, and we also achieved a statistical detection of the host in the remaining 14 from a stack analysis. The UV luminosities can be interpreted in three ways: either as contribution from a passively evolving population of relatively young stars, forming typically 0.5~Gyrs ago, as a mix between a population of old (e.g.\\ $t_{age}\\sim t_\\mathrm{Hubble}$) stars and a small contribution of a recently formed young population (e.g.\\ 0.1\\%--10\\% in mass at an age of 100~Myrs or 1/10th of this for age 10~Myrs), or as an indicator of ongoing star formation at a level of $\\sim$2--15\\,$\\mathrm{M}_\\odot\\:\\mathrm{yr}^{-1}$ (uncorrected for internal dust attenuation). While the first possibility is very simplistic and appears unphysical, the UV colours actually favour the two burst interpretations; but the possibility of on-going star formation cannot be completely ruled out from our data. In the framework of combined old and young populations, it is remarkable how similar the host galaxy colours are within the sample, and, unless different mass--age combinations conspire, hence the estimated stellar mass fractions and ages. The derived young population mass fractions and ages are also very similar to the values estimated in our companion \\gems\\ study of AGN at $z\\la 1$ \\citep{sanc04a}, where we find abnormally blue rest-frame $U-V$ colours for a substantial fraction of host galaxies, particularly the most luminous AGN in the sample. Even more, these colours and ages are in turn very close to the mean values obtained from our ground-based low-$z$ multicolour sample \\citep{jahn04a}. While the stellar population diagnostics of \\citet{kauf03} are not immediately convertible into our simple colour indicators, their impressive and highly significant results point in exactly the same direction. While the results from all these redshift regimes are similar and point to a connection of nuclear activity and the presence of young stars, and that mass fractions of young stars are similar, we always find that a larger range of absolute masses is involved, showing as a range in UV luminosity. Here we find a variation of a factor of $\\ga$40 for a given nuclear luminosity. Our host galaxy colours span a range that reaches the colours of Lyman Break Galaxies for a few very luminous hosts, while, as mentioned, the colours of the majority are somewhat redder than these. A comparison of optical/UV properties to the general population of high redshift galaxies would be very illuminating, but large statistical samples only become available in the near future, e.g.\\ from the \\goods\\ project. This persistent trend to find AGN to be associated with blue stellar colours is intriguing and suggests a close connection between enhanced star formation and nuclear activity. Additional support for such a connection comes from the detection of submm CO emission in a number of extremely luminous high-redshift QSOs and radio galaxies \\citep{omon03}, although the current sensitivity of submm telescopes is insufficient to perform this test for less luminous AGN at high $z$. While the fact that there is a relation can hardly be denied, its physical origin remains obscure. Is the enhancement of star formation a prerequisite for nuclear activity? Is it a simultaneously occurring phenomenon, caused by the same trigger? Or is it a consequence of the AGN? Galaxy merging and interaction are clearly two possible candidates to connect these two phenomena, but neither the only ones nor are the involved physics understood. Much additional data will be required, in particular those helping reliably to reconstruct the star formation history in high-redshift galaxies, before any firm conclusions can be drawn." }, "0403/astro-ph0403181_arXiv.txt": { "abstract": "We revisit the calculation of the abundance of primordial black holes (PBHs) formed from primordial density perturbations, using a formation criterion derived by Shibata and Sasaki which refers to a metric perturbation variable rather than the usual density contrast. We implement a derivation of the PBH abundance which uses peaks theory, and compare it to the standard calculation based on a Press--Schechter-like approach. We find that the two are in reasonable agreement if the Press--Schechter threshold is in the range $\\Dc \\simeq 0.3$ to $0.5$, but advocate use of the peaks theory expression which is based on a sounder theoretical footing. ", "introduction": "Primordial black holes (PBHs) may have formed during the early Universe, and if so can have observational implications at the present epoch, either from effects of their Hawking evaporation or from a contribution to the present dark matter density \\cite{carr,lims}. That there is no unambiguous observational evidence of PBHs is a significant constraint on some possible types of early Universe physics. In particular, they are the only known way of constraining the density perturbation spectrum on extremely short scales, and indeed until fairly recently provided the most powerful upper limit on the spectral index of perturbations with an exactly power-law power spectrum. However, the abundance of PBHs formed from a given initial power spectrum remains uncertain. The traditional calculation takes the same form as the Press-Schechter calculation much used in large-scale structure studies \\cite{PS}, where the density field is smoothed on a mass scale $M$ (in this application taken to be at the time of horizon crossing), and those regions where the density contrast exceeds a threshold value $\\Dc$ are assumed to form PBHs with mass greater than $M$. However the correct value for the threshold is quite uncertain. The `standard' value of 1/3 for a radiation-dominated Universe was derived by Carr \\cite{carr} (see also Ref.~\\cite{harr}), but was probably only ever intended as an order-of-magnitude estimate. Subsequently, Niemeyer and Jedamzik~\\cite{nj} carried out numerical simulations of the collapse of isolated regions and found the threshold for PBH formation, in terms of the relative excess mass within the horizon, to be $\\Delta M / M_{{\\rm h}}= 0.7$. However, Shibata and Sasaki~\\cite{ss} have pointed out that they formulate their initial data after horizon crossing, and hence their criterion cannot be related to the initial perturbations produced by, for instance, a period of inflation. More recently, Shibata and Sasaki~\\cite{ss} devised a new approach to the formation of individual PBHs, seeking to find criteria on the metric perturbation rather than the density field, and in a form which can be applied to superhorizon initial perturbations. They were able to specify a criterion in terms of whether the initial central value of a particular metric perturbation variable $\\psi$ exceeds a threshold value. In this paper, we investigate the implications of this result for the abundance of PBHs formed. ", "conclusions": "We have provided a new calculation of the abundance of PBHs generated by primordial density perturbations. By using a metric perturbation variable rather than the density contrast, a PBH formation criterion can be applied directly to the initial perturbation spectrum. Within this formalism, we have found that the PBH mass spectrum is best computed using the theory of peaks, rather than the standard Press--Schechter-like calculation. Given the considerable uncertainties involved, our results do not lead to any drastic revision of the PBH formation rate, but do put the calculation on a sounder theoretical footing. Our mass function can be fairly well approximated by that of the standard calculation in the region of interest ($\\Omega_{{\\rm PBH}} \\sim 10^{-20}$), if the threshold density $\\Dc$ is taken in the range $0.3$ to $0.5$. This range of threshold values is however significantly lower than the value $\\Dc \\simeq 0.7$ suggested by the simulations of Niemeyer and Jedamzik \\cite{nj}, and in fact encompasses the value $\\Dc = 1/3$ used in the earliest PBH literature. However, we advocate that anyone using our results adopts the peaks theory expression for the mass function given by Eq.~(\\ref{omegapk}). \\vspace*{10pt} After completion of this paper Ref.~\\cite{yok} was brought to our attention. This paper uses the constraints on the metric perturbation variable $\\psi$ from Ref.~\\cite{ss} to calculate the PBH abundance, but does not use the peaks formalism." }, "0403/astro-ph0403339_arXiv.txt": { "abstract": "s{ A sample of radio-loud active galactic nuclei (AGN) at 2cm is studied to test the isotropic distribution of radio sources in the sky and their uniform distribution in space. The sample is complete flux-density limits of 1.5Jy for positive declinations and 2Jy for declinations between $0^{\\circ}<\\delta< -20^{\\circ}$. The active galactic nuclei sample comprises of 133 members. Application of the two-dimensional Kolmogorov-Smirnov test shows that there is no significant deviation from the isotropic distribution in the sky, while the generalised $V/V_{\\rm m}$ test shows that the space distribution of AGN is not uniform at high confidence level (99.9\\%). This is indicative of a strong positive evolution of AGN with cosmic epoch implying that AGN (or jet activity phenomena) were more populous at high redshifts. It is shown that the evolution depends strongly on luminosity: low-luminosity QSOs show a strong positive evolution, while high-luminosity counterparts do not seem to show any evolution with cosmic epoch.} ", "introduction": " ", "conclusions": "" }, "0403/astro-ph0403613_arXiv.txt": { "abstract": "A number of methods of flare prediction rely on classification of physical characteristics of an active region, in particular optical classification of sunspots, and historical rates of flaring for a given classification. However these methods largely ignore the number of flares the active region has already produced, in particular the number of small events. The past history of occurrence of flares (of all sizes) is an important indicator to future flare production. We present a Bayesian approach to flare prediction, which uses the flaring record of an active region together with phenomenological rules of flare statistics to refine an initial prediction for the occurrence of a big flare during a subsequent period of time. The initial prediction is assumed to come from one of the extant methods of flare prediction. The theory of the method is outlined, and simulations are presented to show how the refinement step of the method works in practice. ", "introduction": "Solar flares influence local `space weather,' and as a result there is a demand for accurate flare prediction. Unfortunately no reliable deterministic method of predicting a flare is known, and existing methods are probabilistic in nature. A number of methods discussed in the literature are based on a commonly used white-light classification of sunspots, and the correlation between classification and flare occurrence. The McIntosh classification (McIntosh 1990) categorizes a group of sunspots into one of 60 classes, based on three parameters. Historical flare rates for each of the classifications were used by McIntosh (1990) as the basis of an `expert system' for flare prediction. The system, called Theophrastus (the associated code is called THEO), also incorporates additional information including dynamical properties of spot growth, rotation and shear, magnetic topology inferred from sunspot structure, magnetic classification, and previous flare activity. The method is apparently somewhat subjective, involving rules of thumb incorporated by a human expert. A second approach using the McIntosh classification was presented by Bornmann and Shaw (1994). In this case multiple linear regression was used to determine the effective contribution of each of the McIntosh parameters to the rate of flaring, based on historical records of flaring. Codes based on the methods of McIntosh (1990) and Bornmann and Shaw (1994) are used by the Ionospheric Prediction Service (IPS) of Australia to issue flare predictions at their Learmonth and Culgoora observatories.\\footnote{See http://www.ips.gov.au.} Recently Gallagher, Moon and Wang (2002) implemented a system using historical averages of flare numbers for McIntosh classifications to predict a rate for an active region, and then converted this to a probability of flaring in a day using the assumption of Poisson statistics. This prediction is given as part of the Big Bear Solar Observatory Active Region Monitor (ARM).\\footnote{See http://beauty.nascom.nasa.gov/arm/latest/.} Finally the US National Oceanic and Atmospheric Administration (NOAA) issues flare probability forecasts for active regions which include input from THEO.\\footnote{See http://www.sec.noaa.gov/ftpdir/latest/daypre.txt.} A shortcoming of methods relying on correlations of flaring with active region classification based on historical records is that they ignore the important information of how many flares the active region of interest has already produced. The system of McIntosh (1990) incorporates information about previous activity, but it is unclear how objectively this is done, and the information is limited to the number of large flares already produced by the given active region. In the flare prediction literature, the tendency of a region which has produced large flares in the past to produce large flares in the future is called persistence, which is recognised as one of the most reliable predictors for large flare occurrence in 24-hour forecasts (e.g.\\ Neidig, Weiborg, \\& Seagraves 1989). In this paper we argue that the history of occurrence of all flares (large and small) observed in a given active region is an important indicator as to how the region will flare in the future, and should be used in any prediction. A related criticism of methods based on classification and historical records is that a given classification may embrace active regions with a variety of flaring rates. If an active region has a flaring rate differing from the average historical rate for its class then the predictions will be in error. Studies of solar flare statistics provide simple phenomenological rules describing flare occurrence. It is well known that flares follow a power-law size distribution, where by size we mean e.g.\\ peak flux in soft X-ray. More formally the flare frequency-size distribution $N(S)$ (i.e.\\ the number of events per unit size $S$ and per unit time) may be written \\begin{equation}\\label{eq:pldist} N(S)=AS^{-\\gamma} \\end{equation} where $A$ and $\\gamma$ are constants. The exact power-law index $\\gamma$ depends on the choice of the quantity $S$, but typically it is found to be in the range 1.5 to 2 (e.g.\\ Crosby, Aschwanden, \\& Dennis 1992). The power law index $\\gamma$ appears to be the same in different active regions~\\cite{whe00}, although there is some evidence that it varies with the solar cycle~\\cite{bai93}. A second simple rule concerns the way flares occur in time. Studies of the rate of occurrence of soft X-ray flares in individual active regions suggest that events occur as a Poisson process in time (e.g.\\ Moon et al.\\ 2001), although many active regions exhibit changes in the mean rate of events (Wheatland 2001). In this paper we show how the observed record of flaring in an active region may be used together with the phenomenological rules of flare statistics to objectively refine an initial flare prediction. The initial prediction may be based on the McIntosh classification, or may come from any other prediction method which does not consider the flare data. The new method is envisaged to work as follows. When an active region appears at the east limb of the Sun, the best guess as to its future flare productivity comes from one of the conventional prediction methods. However, as the active region produces flares, the observed flare statistics are used to adjust the prediction for future flaring. After many flares have been observed, the prediction for future flaring may be dominated by the contribution from the observed data. This process --- refining a probability estimate based on new data --- is naturally performed using Bayes's theorem (e.g.\\ Sivia 1996; Jaynes 2003). The layout of the paper is as follows. In \\S\\,2 a simple approach to flare prediction using only the past record of flaring from an active region [previously presented in Wheatland (2001)] is reiterated. In \\S\\,3 the new method of prediction, combining existing methods and information from observed flare statistics, is described. In \\S\\,4 simulations are presented showing how the method uses the observed flaring record, and in \\S\\,5 the results are discussed. ", "conclusions": "Existing methods of solar flare prediction do not make complete use of an important source of information: the time history of flares already observed in the active region of interest, in particular frequently occurring small events. A new method for flare prediction is presented which exploits the observed history of flaring from an active region to improve an initial prediction, which e.g.\\ may come from one of the existing methods. To make the example concrete we may think of the initial prediction coming from from the McIntosh sunspot classification, which is a common basis for prediction. This background information provides an initial estimate for the expected flaring rate through a prior distribution $\\Lambda_{\\rm MC}(\\lambda_1)$, which represents the probability that the flaring rate above a (small) size $S_1$ is $\\lambda_1$, given historical rates of occurrence of flares for the given McIntosh class. Bayes's theorem is then used to estimate the probability $\\epsilon$ of observing a large flare (above size $S_2$) in a given period of time, based on this prior information and on the sequence of flares already produced by the active region, and assuming simple phenomenological rules describing the occurrence of flares. In this paper the basic theory behind the inference of $\\epsilon$ based on observed data is presented. The inclusion of background information [i.e.\\ the construction of the priors $\\Lambda_{\\rm MC}(\\lambda_1)$] is yet to be done. The method relies on event sizes following the phenomenological law~(\\ref{eq:pldist}). Some studies of very small extreme ultraviolet events (`nanoflares') suggest that their thermal energies follow a steeper distribution than energies of large events (e.g.\\ Krucker and Benz 1998; Parnell and Jupp 2000), although this remains controversial (e.g.\\ Aschwanden and Parnell 2002). From the point of view of the prediction method presented here, the uncertainty over the low-size end of the distribution is irrelevant provided events significantly larger than nanoflares are used. In any case the observed distributions from many active regions may be examined as a check on Equation~(\\ref{eq:pldist}). A related point is that the distribution~(\\ref{eq:pldist}) requires a cutoff at large sizes on energetics grounds, and neglect of this cutoff will lead to the number of large flares being overestimated. A cutoff will be incorporated before the method is applied to real data. The choice of the quantity $S$ has not been addressed, although a good choice is likely to be important to the method. Most flare forecasting deals with soft X-ray events, in particular prediction of GOES (Geostationary Observational Environmental Satellite) M and X class events (events with peak fluxes greater than $10^{-5}$W/m$^2$ and $10^{-4}$W/m$^2$ respectively in the 1-8 Angstrom band observed by the satellites). A practical motivation for this is that flare soft X-ray emission causes disturbances of the ionosphere which affect shortwave radio communication, and there is a need to predict these occurrences. A disadvantage of using GOES events is that they are not ideal for flare statistics e.g.\\ because of problems with event selection due to the large background in soft X-ray (see Wheatland 2001). A number of other issues also need to be considered before the method is implemented with real data. A point neglected so far is that active regions evolve, so that predictions based on the traditional methods also change with time. For example, an active region evolves through McIntosh classifications (e.g.\\ Bornmann, Kalmbach, Kulhanek, and Casale 1990). Changes in background information such as this should be incorporated through changes in the prior, and this question will be considered in more detail in future work. A related point concerns the construction of the prior distributions for rate. It is likely that the McIntosh classification will be used, although other possibilities will be considered. The problem is then to determine the probability of a given McIntosh class having a given rate, based on observed flaring sequences in the historical record for active regions of that class. The details of this calculation will be addressed in future work. Finally, as with all methods of forecasting, it is essential to test the reliability of the method. It is straightforward to compare, after the fact, the number of predicted and the number of observed events for a large sample of active regions. The method presented here will be implemented and tested in this way, and the results compared with existing methods of prediction." }, "0403/astro-ph0403425_arXiv.txt": { "abstract": "We propose a dynamical mechanism for capturing stars around a massive black hole (MBH), which is based on the accumulation of a very dense cluster of compact stellar remnants near the MBH. This study is motivated by the presence of $\\sim\\!10$ young massive stars ($\\Ms\\!\\sim\\!3$--$15\\,\\Mo$, spectral types $\\sim$B9V--O8V) less than $0.04$ pc from the MBH in the Galactic center (GC). Their existence in the extreme environment so close to an MBH is a challenge for theories of star formation and stellar dynamics. We show that young stars, which formed far from the MBH and were then scattered into eccentric orbits, repeatedly cross a cluster of stellar black holes (SBHs), where they may undergo rare direct three-body exchanges with an MBH-SBH {}``binary''. The interaction between two objects of comparable mass ejects the SBH and captures the star on a tight orbit around the MBH. Such captures can naturally explain some trends observed in the orbits of the young stars. We derive the capture cross-section, validate it by Monte Carlo simulations, and calculate the number of captured stars in the GC using the currently uncertain estimates of the numbers of SBHs in the inner 0.04 pc and of young stars in the inner few parsecs of the GC. We find that under favorable conditions three-body exchange can account for $\\sim\\!25$\\% of the observed stars, mostly at the fainter end of the observed range. We discuss additional effects that possibly increase the capture efficiency. Future detections of the dark mass around the MBH and deeper surveys of the central parsecs will establish whether or not there are enough SBHs and young stars there for exchange captures to singly account for the central young stars. We estimate that there are also $\\sim\\!35$ lower mass stars ($\\Ms\\!\\sim\\!1$--$3\\,\\Mo$, $\\sim$G2V--A0V) in the inner $0.04$ pc similarly captured by exchanges with neutron stars (NSs). Ongoing replacement of compact remnants by main-sequence stars (SBHs by NS progenitors, NSs by white dwarf progenitors) may regulate the accumulation of compact remnants near the MBH. ", "introduction": "Deep near-infrared photometric (Krabbe et al. \\citeyear{Kra95}; Genzel et al. \\citeyear{Gen03a}), spectroscopic (Genzel et al. \\citeyear{Gen97}; Eckart, Ott \\& Genzel \\citeyear{Eck99}; Figer et al. \\citeyear{Fig00}; Gezari et al. \\citeyear{Gez02}; Ghez et al. \\citeyear{Ghe03a}) and astrometric (Ghez et al. \\citeyear{Ghe03b}; Sch\\\"odel et al. \\citeyear{Sch03}) observations of the dense stellar cusp around the massive black hole (MBH) in the Galactic center (GC) reveal a centrally concentrated distribution of young massive stars, $\\Ns\\sim\\!10$ stars within the central $\\rs\\sim\\!0.04$ pc, $\\sim\\!40$ within $\\sim\\!0.1$ pc, spanning a mass range of $\\sim\\!3$--$15\\,\\Mo$ (spectral types $\\sim$B9V--O8V) with a median mass of $\\Ms\\!\\sim\\!10\\,\\Mo$, radius of $\\Rs\\!\\sim\\!4.5\\,\\Ro$ and main sequence life time of $\\ts\\!\\sim\\!3\\!\\times\\!10^{7}$ yr% \\footnote{These are rough inferences. The separation of the young stars from the old population is uncertain since at present only the brightest star has been spectroscopically identified (Ghez et al. \\citeyear{Ghe03a}).% }. Orbital solutions obtained for eight of the stars (Ghez et al. \\citeyear{Ghe03b}; Sch\\\"odel et al. \\citeyear{Sch03} ) tentatively suggest some trends in their orbital properties: a lower bound on the apoapse of $\\sim\\!0.01$ pc (Ghez et al. \\citeyear{Ghe03b}) and higher than random orbital eccentricities (Sch\\\"odel et al. \\citeyear{Sch03}). None of the solutions proposed so far for the puzzle of the young stars (Genzel et al. \\citeyear{Gen03a}; Ghez et al. \\citeyear{Ghe03a}) are satisfactory. These fall into three categories: exotic modes of star formation near the MBH; rejuvenation of old stars in the local population; or dynamic migration from farther out, where stars can form. Even if shock cooling of molecular gas by cloud-cloud collisions could trigger star formation near the MBH (Morris \\citeyear{Mor93}), molecular clouds would either form stars or be tidally disrupted well outside of the inner 0.1 pc (Vollmer \\& Duschl \\citeyear{Vol01}). Growth and rejuvenation by mergers (Genzel et al. \\citeyear{Gen03a}) are not expected to be efficient in high velocity collisions near an MBH. Tidal heating by the MBH requires that the stars approach the MBH much closer than they are observed to do (Alexander \\& Morris \\citeyear{Ale03a}). The young stars are too short-lived to have formed far from the MBH and then migrated inward by mass segregation or dynamical friction. The migration can be accelerated if the stars are associated with a massive {}``anchor'': an extremely dense young cluster (Portegies Zwart, McMillan \\& Gerhard \\citeyear{Por03}; Kim \\& Morris \\citeyear{Kim03}), a very massive binary companion (Gould \\& Quillen \\citeyear{Gou03}), or a $10^{3}$--$10^{4}\\,\\Mo$ black hole (Hansen \\& Milosavljevi\\'c \\citeyear{Han03}). However, these scenarios must assume the existence of very rare, or even hypothetical objects, or else they cannot bring the stars close enough to the MBH. Such processes may possibly explain the separate population of very massive and luminous {}``He stars'' $0.1$--$0.5$ pc from the MBH (Krabbe et al. \\citeyear{Kra95}), which we do not attempt to model here. Our model is based on the fact that $10^{4}$--$10^{5}$ stellar black holes (SBHs) of mass $\\sim\\!7$--$10\\,\\Mo$ are estimated to exist within $\\sim\\!1$ pc of the MBH in the GC, where they have been accumulating by dynamical friction over the lifetime of the Galaxy ($t_{H}\\!\\sim\\!10\\,\\mathrm{Gyr}$ ) from a {}``collection basin'' $\\sim\\!10$ pc wide (Morris \\citeyear{Mor93}; Miralda-Escud\\'{e} \\& Gould \\citeyear{Mir00}). Numerical simulations of the evolution of the GC (Freitag \\citeyear{Fre03}) confirm that the SBHs sink to the center on a short timescale of a few gigayears, settle into a centrally concentrated distribution where the enclosed number scales as $\\ns(<\\! r)\\!\\propto r^{5/4}$ (Bahcall \\& Wolf \\citeyear{Bah77}), and dominate the stellar mass there. ", "conclusions": "The efficiency of a {}``billiard ball'' recoil depends strongly on the mass ratio of the colliding objects. The dynamical evolution of a stellar system around an MBH naturally provides a dense concentration of targets whose masses are well matched for stopping and capturing unbound young stars like those observed very near the MBH in the GC. The attempt to predict the number of captured stars is limited by the uncertainty in the mass and number distribution of the SBHs and in the number and orbital properties of young stars (spectral types $\\sim$B9V--O8V) in the inner few parsecs of the GC. We show that under favorable conditions, capture can account for $\\sim\\!25$\\% of the observed young stars. Additional effects that were not taken into account here may increase the capture efficiency. We considered only capture by a single direct exchange in the point-mass approximation. However, a few weaker interactions may also lead gradually to capture during the star's lifetime. Further study is needed to estimate the contribution of multiple scatterings to the capture cross section. Tidal energy extraction in captures near the tidal limit also increases the capture cross section there. A possible outcome of internal mixing by a strong tidal interaction is an extended main-sequence lifetime and higher luminosity (Maeder \\& Meynet \\citeyear{Mae00}). This would further increase the number of captured luminous stars (Eq. \\ref{eq:iNs}). In the context of direct exchange capture, the stars with the smallest apoapse offer an opportunity to study the long term effects of a strong tidal interaction. Interestingly, S2, the star with the smallest apoapse, is also the brightest (Ghez et al. \\citeyear{Ghe03b}; Sch\\\"odel et al. \\citeyear{Sch03}). The orbital eccentricity distribution reflects the mass ratio between the star and the SBH. Thus, a spectral determination of the masses of the captured stars together with a statistical analysis of their eccentricities can probe the poorly known mass function of SBHs. The total number of captured stars over the lifetime of the Galaxy (assuming steady state), $\\sim\\!2\\Ns t_{H}/\\ts\\!\\sim\\!3\\!\\times\\!10^{4}$ (for $40$ stars of $10\\,\\Mo$ in 0.1 pc), is of the same order as the number of SBHs in the inner parsec. Thus, if the dynamical friction timescale at $\\left\\langle a_{0}\\right\\rangle $ (where the SBHs are ejected to) is not much smaller than the age of the Galaxy, the continual replacement of SBHs by NS progenitors and of NSs by white dwarf progenitors (Fryer \\& Kalogera \\citeyear{Fry01}) may regulate the build-up of the dense cusp of compact stellar remnants." }, "0403/astro-ph0403080_arXiv.txt": { "abstract": "{We present photometry and spectroscopy of the afterglow of \\grb. VLT spectra of the afterglow show damped \\lya\\ (DLA) absorption and low- and high-ionization lines at a redshift $z$=3.3718$\\pm$0.0005. The inferred neutral hydrogen column density, log N(\\ion{H}{i})=21.90$\\pm$0.07, is larger than any (GRB- or QSO-) DLA \\ion{H}{i} column density inferred directly from \\lya\\ in absorption. From the afterglow photometry, we derive a conservative upper limit to the host-galaxy extinction: A$_{\\rm V}$$<$0.5 mag. The iron abundance is [Fe/H]=--1.47$\\pm$0.11, while the metallicity of the gas as measured from sulphur is [S/H]=--1.26$\\pm$0.20. We derive an upper limit on the H$_2$ molecular fraction of 2$N$(H$_2$)/(2$N$(H$_2$)+$N$(\\ion{H}{i}))$\\lsim$10$^{-6}$. In the \\lya\\ trough, a \\lya\\ emission line is detected, which corresponds to a star-formation rate (not corrected for dust extinction) of roughly 1 M\\subsun\\ yr$^{-1}$. All these results are consistent with the host galaxy of \\grb\\ consisting of a low metallicity gas with a low dust content. We detect fine-structure lines of silicon, \\ion{Si}{ii}*, which have never been clearly detected in QSO-DLAs; this suggests that these lines are produced in the vicinity of the GRB explosion site. Under the assumption that these fine-structure levels are populated by particle collisions, we estimate the \\ion{H}{i} volume density to be n$_{\\ion{H}{i}}=10^2-10^4$ cm$^{-3}$. HST/ACS imaging 4 months after the burst shows an extended AB(F606W)=28.0$\\pm$0.3 mag object at a distance of 0\\farcs14 (1kpc) from the early afterglow location, which presumably is the host galaxy of \\grb. ", "introduction": "\\label{sec:introduction} Damped \\lya\\ (DLA) absorbers, conventionally detected in Quasi-Stellar Object (QSO) spectra, are absorption-line systems that have a column density of N(\\ion{H}{i}) $\\ge$ 2$\\times 10^{20}$ atoms cm$^{-2}$, as determined from the damping wings of the \\lya\\ line \\citep[e.g.][]{1986ApJS...61..249W,1989ApJ...344..567T}. DLA systems are believed to contain the bulk of the neutral hydrogen at high redshift and to be the gas reservoir from which the stars at the present epoch are produced \\citep[e.g.][]{1987txra.symp..309W,1991ApJS...77....1L}. Numerous high-resolution spectroscopic studies have extracted detailed information about the metallicity \\citep[e.g.][]{prochaska}, the kinematics \\citep{1997ApJ...487...73P,1998A&A...337...51L}, and the dust and H$_2$ contents \\citep{2000A&A...364L..26P,ledoux} of DLA systems as a function of redshift. Despite intensive searches, only a handful of DLA counterparts have been detected so far \\citep[see][]{2002ApJ...574...51M}; linking DLA systems with galaxy types has therefore proven difficult: some advocate large, disk-forming galaxies \\citep[e.g.][]{1995ApJ...454..698W,1997ApJ...487...73P}, others suggest they are faint, gas-rich dwarfs \\citep{1998ApJ...495..647H}. Gamma-ray burst (GRB) afterglows are, just as QSOs, bright and distant sources. For instance, the spectacular GRB\\,990123 was detected at the 9$^{\\rm th}$ visual magnitude \\citep{1999Natur.398..400A} while it was located at $z$=1.6 \\citep{1999Natur.398..389K,1999Sci...283.2075A}. However, the afterglow brightness in general fades very rapidly in time (roughly flux $\\propto$ time$^{-1}$). The current afterglow redshifts range from $z$=0.169 \\citep{2003GCN..2020....1G} to $z$=4.5 \\citep{2000A&A...364L..54A}. Moreover, GRBs are associated with massive-star formation: the discovery of a supernova (SN) spectrum similar to that of SN1998\\,bw \\citep{1998Natur.395..670G} superimposed on the GRB\\,030329 afterglow spectrum \\citep{stanek,hjorth030329} provided strong evidence that at least some of the long-duration ($\\gsim$ 2s) GRBs are caused by the core collapse of massive stars \\citep{1993ApJ...405..273W,1999ApJ...524..262M}. The discovery of a damped \\lya\\ (DLA) absorption line at the burst redshift in the spectra of several GRB afterglows \\citep{2001A&A...370..909J,fynbo926spectrum,hjorth020124} is consistent with the massive-star progenitor scenario: they indicate a high neutral hydrogen column density origin in the host galaxy, presumably a star-forming region. However, the signal-to-noise ratio at the location of the DLA absorption line in the spectra is fairly low in these cases, much lower than for typical QSO-DLAs. We here present afterglow spectra of the high-redshift \\grb, which unambiguously demonstrate a GRB-DLA, with a column density exceeding that of any (QSO- or GRB-) DLA measured so far using \\lya\\ in absorption. These spectra allow us to measure the metallicity of the host environment and obtain an upper limit on the molecular fraction, i.e. measurements that are routinely performed for QSO-DLAs, but that are still unique for GRB hosts. Although the GRB-DLA sample is still very small, we compare them with QSO-DLAs in two aspects: their \\ion{H}{i} column density and their metallicity. \\grb\\ was detected on 23 March 2003 at 21:57 UT by HETE \\citep{2003GCN..1956....1G} with a fluence of 1.1$\\times$10$^{-6}$ ergs cm$^{-2}$ (30-400 keV), and a duration of 26 seconds. Following the HETE localization, the optical counterpart was discovered 7.6 hours after the burst at R.A. 11\\hour06\\min09\\fs38, Decl. --21\\degree46\\arcmin13\\farcs3 (J2000) by \\citet{2003GCN..1949....1G}, with a brightness of R=18.7. Our team reported a preliminary redshift of $z$=3.372 \\citep{2003GCN..1953....1V}, which is currently the third highest redshift for a GRB \\citep{1998Natur.393...35K,2000A&A...364L..54A}. This paper is organized as follows: in Sect. \\ref{sec:observations}, we describe the data reduction of both the spectroscopic and imaging observations. In Sect. \\ref{sec:photometry}, we present the light curves and infer an upper limit on the rest-frame optical extinction. We measure the equivalent widths of the absorption lines and determine the burst redshift in Sect. \\ref{sec:redshift}. An \\ion{H}{i} column density model is fitted to the damped \\lya\\ line in Sect. \\ref{sec:column}, and we analyze the spectra in more detail in Sect. \\ref{sec:metallicityandh2} to derive the metallicity and an upper limit on the molecular hydrogen (H$_2$) fraction. The detection of \\lya\\ in emission is presented in Sect. \\ref{sec:lya}, and we report on the detection of the probable host galaxy of \\grb\\ in HST/ACS imaging data in Sect. \\ref{sec:hst}. In the final section, we close with a general discussion of all these results. \\begin{table}[tp] \\centering \\caption[]{Log of UT4/FORS2 spectroscopic observations}\\label{tab:spectroscopy} \\null\\vspace{-1.0cm} $$ \\begin{array}{cllccc} \\hline \\noalign{\\smallskip} \\rm UT \\, date & \\rm grism(filter) & \\rm coverage & \\rm \\lambda / \\Delta\\lambda & \\rm exptime & \\rm seeing \\\\ \\rm March\\,03 & & \\rm (nm) & & \\rm (min) & (\\arcsec) \\\\ \\hline \\rm 25.050 & \\rm 300V & 330-660 & 440 & 3\\times10 & 1.1 \\\\ \\rm 25.077 & \\rm 300I(OG590) & 600-1100 & 660 & 3\\times10 & 0.7 \\\\ \\rm 26.213 & \\rm 1400V & 456-586 & 2100 & 4\\times30 & 0.8 \\\\ \\rm 26.306 & \\rm 1200R(GG435) & 575-731 & 2140 & 4\\times30 & 0.9 \\\\ \\hline \\end{array} $$ \\end{table} \\begin{table}[tbp] \\centering \\caption[]{Log of imaging observations}\\label{tab:imaging} \\null\\vspace{-1.35cm} $$ \\begin{array}{rccccc} \\hline \\noalign{\\smallskip} \\rm UT \\, date & \\rm magnitude ^{\\mathrm{a}} & \\rm filter & \\rm exptime & \\rm seeing & \\rm tel./instr.^{\\mathrm{b}} \\\\ \\rm (2003) & & & \\rm (min) & (\\arcsec) & \\\\ \\hline \\rm Mar\\, 24.302 & 20.39 \\pm 0.06 & \\rm B & 12 & 2.4 & \\rm USNO\\,1m \\\\ \\rm Mar\\, 24.310 & 19.68 \\pm 0.05 & \\rm V & 8 & 2.5 & \\rm USNO\\,1m \\\\ \\rm Mar\\, 24.316 & 18.75 \\pm 0.03 & \\rm R & 8 & 2.6 & \\rm USNO\\,1m \\\\ \\rm Mar\\, 24.323 & 18.20 \\pm 0.04 & \\rm I & 8 & 2.4 & \\rm USNO\\,1m \\\\ \\rm Mar\\, 24.992 & 20.64 \\pm 0.26 & \\rm R &8.33 & 1.1 & \\rm CAHA\\,2.2m \\\\ \\rm Mar\\, 25.027 & 20.56 \\pm 0.08 & \\rm R & 54 & 2.0 & \\rm Danish \\\\ \\rm Mar\\, 25.030 & 21.44 \\pm 0.03 & \\rm V & 1 & 1.1 & \\rm FORS2 \\\\ \\rm Mar\\, 25.033 & 21.45 \\pm 0.03 & \\rm V & 1 & 1.0 & \\rm FORS2 \\\\ \\rm Mar\\, 25.091 & 22.29 \\pm 0.04 & \\rm B & 1 & 0.9 & \\rm FORS2 \\\\ \\rm Mar\\, 25.092 & 21.42 \\pm 0.02 & \\rm V & 1 & 0.8 & \\rm FORS2 \\\\ \\rm Mar\\, 25.094 & 20.51 \\pm 0.02 & \\rm R & 1 & 0.8 & \\rm FORS2 \\\\ \\rm Mar\\, 25.095 & 20.02 \\pm 0.02 & \\rm I & 1 & 0.8 & \\rm FORS2 \\\\ \\rm Mar\\, 25.099 & 20.04 \\pm 0.10 & \\rm I & 15 & 2.0 & \\rm Danish \\\\ \\rm Mar\\, 25.129 & 20.57 \\pm 0.03 & \\rm R & 25 & 2.0 & \\rm Danish \\\\ \\rm Mar\\, 25.150 & 21.55 \\pm 0.12 & \\rm V & 15 & 1.8 & \\rm Danish \\\\ \\rm Mar\\, 25.250 & 20.83 \\pm 0.13 & \\rm R & 130 & 3.0 & \\rm SARA \\\\ \\rm Mar\\, 26.023 & 22.17 \\pm 0.08 & \\rm V & 25 & 1.2 & \\rm Danish \\\\ \\rm Mar\\, 26.041 & 21.27 \\pm 0.05 & \\rm R & 25 & 1.0 & \\rm Danish \\\\ \\rm Mar\\, 26.064 & 20.82 \\pm 0.08 & \\rm I & 25 & 1.0 & \\rm Danish \\\\ \\rm Mar\\, 26.107 & 19.86 \\pm 0.04 & \\rm J & 6 & 0.7 & \\rm NACO \\\\ \\rm Mar\\, 26.120 & 17.93 \\pm 0.07 & \\rm K &2.25 & 0.7 & \\rm NACO \\\\ \\rm Mar\\, 26.162 & 22.12 \\pm 0.02 & \\rm V & 1 & 0.6 & \\rm FORS2 \\\\ \\rm Mar\\, 26.164 & 22.09 \\pm 0.03 & \\rm V & 1 & 0.6 & \\rm FORS2 \\\\ \\rm Mar\\, 26.257 & 22.22 \\pm 0.03 & \\rm V & 1 & 0.7 & \\rm FORS2 \\\\ \\rm Mar\\, 26.259 & 22.18 \\pm 0.04 & \\rm V & 1 & 0.8 & \\rm FORS2 \\\\ \\rm Mar\\, 26.267 & 21.50 \\pm 0.07 & \\rm R & 160 & 2.2 & \\rm USNO\\,1m \\\\ \\rm Mar\\, 26.306 & 18.13 \\pm 0.18 & \\rm K &3.35 & 0.5 & \\rm UKIRT \\\\ \\rm Mar\\, 26.351 & 22.32 \\pm 0.04 & \\rm V & 3 & 0.8 & \\rm FORS2 \\\\ \\rm Mar\\, 26.351 & 19.38 \\pm 0.05 & \\rm H &3.35 & 0.6 & \\rm UKIRT \\\\ \\rm Mar\\, 26.353 & 21.43 \\pm 0.02 & \\rm R & 3 & 0.8 & \\rm FORS2 \\\\ \\rm Mar\\, 26.356 & 20.90 \\pm 0.03 & \\rm I & 3 & 0.8 & \\rm FORS2 \\\\ \\rm Mar\\, 26.379 & 20.09 \\pm 0.03 & \\rm J &3.35 & 0.7 & \\rm UKIRT \\\\ \\rm Mar\\, 27.041 & 22.01 \\pm 0.14 & \\rm R & 30 & 1.0 & \\rm Danish \\\\ \\rm Mar\\, 27.156 & 22.94 \\pm 0.09 & \\rm V & 24 & 0.7 & \\rm Gemini\\,S \\\\ \\rm Mar\\, 27.176 & 22.15 \\pm 0.15 & \\rm R &22.5 & 0.6 & \\rm Gemini\\,S \\\\ \\rm Mar\\, 27.196 & 21.57 \\pm 0.09 & \\rm I &22.5 & 0.6 & \\rm Gemini\\,S \\\\ \\rm Mar\\, 28.159 & 22.36 \\pm 0.09 & \\rm R & 36 & 1.2 & \\rm Danish \\\\ \\rm Mar\\, 28.398 & 19.06 \\pm 0.20 & \\rm K & 7 & 0.5 & \\rm UKIRT \\\\ \\rm Mar\\, 28.455 & 20.34 \\pm 0.08 & \\rm H & 7 & 0.7 & \\rm UKIRT \\\\ \\rm Apr\\, 3.282 & 24.30 \\pm 0.13 & \\rm R & 88 & 1.3 & \\rm Gemini\\,S \\\\ \\rm Jul\\, 5.979 & > 25.0 \\,(3\\sigma) & \\rm I & 15 & 0.7 & \\rm FORS2 \\\\ \\rm Jul\\, 5.993 & > 25.6 \\,(3\\sigma) & \\rm R & 15 & 0.8 & \\rm FORS2 \\\\ \\rm Jul\\, 20.958 & 28.0 \\pm 0.3 & \\rm V & 32 & & \\rm HST \\\\ \\hline \\end{array} $$ \\begin{list}{}{} \\item[$^{\\mathrm{a}}$] The magnitudes have {\\it not} been corrected for Galactic extinction, and the errors do {\\it not} include the uncertainty in the absolute calibration (see text). \\item[$^{\\mathrm{b}}$] Combinations of telescopes and instruments: USNO Flagstaff Station 1.0m with 2k$\\times$2k Tek CCD, 0$\\farcs$68/pixel; Calar Alto 2.2m and CAFOS with 1k$\\times$1k SITe CCD, 0$\\farcs$29/pixel; ESO/Danish 1.54m and DFOSC with 2k$\\times$4k EEV CCD, binned to 0$\\farcs$78/pixel; Yepun and FORS2 with two MIT CCDs of 4k$\\times$2k, binned to 0$\\farcs$25/pixel; Yepun and NACO with the Aladdin InSb 1k$\\times$1k detector and S54 camera, 0$\\farcs$054/pixel; SARA 0.9m and Apogee Ap7 CCD camera with 512$\\times$512 array, 0$\\farcs$7/pixel; UKIRT and UFTI with a 1k$\\times$1k HgCdTe array, 0$\\farcs$091/pixel; Gemini South and acquisition camera with 1K$\\times$1K EEV CCD, binned to 0$\\farcs$23/pixel; HST/ACS and WFC detector with two 2k$\\times$4k SITe CCDs, 0$\\farcs$05/pixel. \\end{list} \\end{table} ", "conclusions": "\\label{sec:discussion} Our observations show that \\grb\\ occurred behind a very high \\ion{H}{i} column density, in an environment (immediate and host-galaxy combined) having a low molecular hydrogen fraction (f$\\lsim$10$^{-6}$), a low metallicity ([S/H]=--1.26$\\pm$0.20) and a low dust content ($\\kappa$=0.02). For the DLA host of GRB\\,020124, \\citet{hjorth020124} also find evidence for a large \\ion{H}{i} column density with a low reddening. The inferred low dust content may be interpreted as a selection bias: GRBs that would occur in a dusty host galaxy would be harder to detect because they would then be fainter. However, in this case one would expect to observe many GRB afterglows with considerable extinction in the optical, for which there is no clear evidence \\citep{2001ApJ...549L.209G,2002MNRAS.330..583L}. In apparent contradiction with this is the detection of several host galaxies in the radio and sub-mm \\citep[e.g.][]{frail222,2003MNRAS.338....1B,2003ApJ...588...99B}, suggesting that at least some GRB hosts are dusty, as expected when most of the star formation in the universe occurs in submm-bright galaxies \\citep[see][]{2002MNRAS.329..465R}. Dust destruction \\citep[e.g.][]{2001ApJ...563..597F,2002ApJ...569..780D,2003ApJ...585..775P}, which has been proposed to account for the apparent discrepancy between the low optical extinction and large (X-ray) gas column densities \\citep{2001ApJ...549L.209G}, could play a role, but is not required by our data: the reduced metallicity and hence the low dust-to-gas ratio in the host of \\grb\\ is sufficient to explain the combination of a large \\ion{H}{i} column density with a low optical extinction \\citep[see also][]{hjorth020124}. \\lya\\ in emission is detected for \\grb, and we inferred a star-formation rate of about 1 M\\subsun\\ yr$^{-1}$, which is in good agreement with the SFR value that we obtained from the UV continuum emission of the host galaxy. \\citet{fynbolya} note that \\lya\\ is commonly observed in all GRB host galaxies at high redshift for which it could be detected. In contrast, only 25\\% of the Lyman-break galaxies are \\lya\\ emitters with an equivalent width EW $>$ 20 \\AA\\ \\citep{2003ApJ...588...65S}. \\citet{fynbolya} suggest that this difference is due to GRB hosts having a low metallicity and a low dust content, consistent with our observations of \\grb\\ and with those of GRB\\,020124 \\citep{hjorth020124}. We note that QSO-DLAs also have a low metallicity and a low dust content, but they rarely show \\lya\\ in emission. However, since most galaxy counterparts of QSO-DLAs are very faint, \\lya\\ in emission is not expected to be detected in most cases with the current detection limits \\citep[see][]{1999MNRAS.305..849F}. The low-dust inference for GRB\\,020124 \\citep{hjorth020124} and \\grb\\ is different from the results of \\citet{2003ApJ...585..638S}, who find evidence for a high dust content in three GRB host galaxies. From the fine-structure lines \\ion{Si}{ii}* \\lam\\lam 1309,1533 , we estimated the \\ion{H}{i} volume density of the gas producing this absorption: n$_{\\ion{H}{i}}=10^2-10^4$ cm$^{-3}$, under the assumption that these fine-structure levels are populated by collisions, and not through direct excitation by infra-red photons (which is not an important excitation mechanism in the case of \\ion{Si}{ii}*), or fluorescence \\citep[see][]{2002MNRAS.329..135S}. This volume density is higher than inferred for QSO-DLA environments \\citep{2002MNRAS.329..135S}, but typical of Galactic molecular clouds \\citep[e.g.][]{1999osps.conf....3B,2002ApJ...565..174R}. As these lines has never been clearly detected up to now in QSO-DLAs, the detection of these {Si}{\\sc ii}* lines in the \\grb\\ spectrum suggests an origin in the vicinity of the GRB place of birth (e.g. the star-forming region in which it exploded). Combining the measured \\ion{H}{i} column density with the order of magnitude estimate of the \\ion{H}{i} volume density, we obtain a size (diameter) of $\\sim$ 5pc (taking n$_{\\ion{H}{i}}=10^3$ cm$^{-3}$) and a mass of $\\sim$ 2$\\times$10$^3$ M\\subsun\\ for the \\ion{Si}{ii}* absorbing region. With the volume density so high, one would expect hydrogen molecules to be present, which, surprisingly, we do not detect. We obtain a rather strong upper limit on the mean molecular fraction of the gas in the GRB environment and the host galaxy: f$\\equiv$2$N$(H$_2$)/(2$N$(H$_2$)+$N$(\\ion{H}{i}))$\\lsim$10$^{-6}$. This could be explained by the low metallicity of the gas \\citep[see][]{ledoux}, but it may also be that the molecules in the GRB environment have been dissociated by the strong GRB UV/X-ray emission \\citep[e.g.][]{2002ApJ...569..780D}. In the latter case, however, the UV/X-ray flash would also ionize the neutral gas in the GRB vicinity \\citep[see][]{2002ApJ...569..780D}, which would make the high \\ion{H}{i} column density detection improbable. Therefore, a large fraction of the \\ion{H}{i} column density may not be located close to the GRB explosion site, but elsewhere in the host galaxy, while the high volume density \\ion{Si}{ii}* region (and the expected molecular hydrogen), is located in the vicinity of the burst. In this case, the disks of GRB host galaxies need to be much denser than the Galactic disk, as 7 random sight lines through the disk toward the location of the Earth would not result in 5 \\ion{H}{i} column densities above 10$^{21}$ cm$^{-2}$ \\citep[see Fig. 5 of][]{1990ARA&A..28..215D}, as is observed for GRB sightlines (see Fig. \\ref{histo}). Finally, the population of the \\ion{Si}{ii}* levels may have been partly caused by fluorescence of photons from the GRB itself, in which case the volume density estimate above is a strict upper limit. The \\ion{H}{i} column density that we inferred toward \\grb\\ is higher than that of any (QSO- or GRB-) DLA measured using \\lya\\ in absorption. It is generally assumed that the apparent \\ion{H}{i} column density limit of N(\\ion{H}{i})$\\sim$10$^{22}$ atoms cm$^{-2}$ for QSO-DLAs is due to an observational bias against the detection of such high-column density systems, as these would obscure the background QSO if they contain some dust \\citep[e.g.][]{1984ApJ...278....1O,1993ApJ...402..479F}. However, a radio-selected QSO survey for DLA systems by \\citet{2001A&A...379..393E} did not uncover a previously unrecognized population of N(\\ion{H}{i})$>10^{21}$ cm$^{-2}$ DLA systems in front of faint QSOs. An alternative scenario was proposed by \\citet{2001ApJ...562L..95S}: the lack of high \\ion{H}{i} column density systems could be due to the conversion of \\ion{H}{i} to H$_2$ as the neutral gas density increases. This picture is consistent with observations of Galactic molecular clouds \\citep[e.g.][]{1999osps.conf....3B}. In \\grb, however, we do not find any evidence for the presence of H$_2$ in addition to \\ion{H}{i} to support this scenario. Future GRBs with possibly even larger \\ion{H}{i} column densities than that toward \\grb\\ could provide further constraints to the existence of a rapid conversion of \\ion{H}{i} to H$_2$ at high \\ion{H}{i} column densities. We compared the metallicities and \\ion{H}{i} column densities of the (still very small) sample of GRB-DLAs with QSO-DLAs, and we found both quantities to be higher in GRB-DLAs than in QSO-DLAs. This is not surprising, as GRBs are now known to probe massive-star forming regions \\citep{stanek,hjorth030329} where the gas density and the metallicity are higher than along random QSO sight lines through foreground galaxies. A KS test applied to the column densities shows that the probability that the GRB- and QSO-DLA samples are drawn from the same parent distribution is very low (0.0006). On the other hand, two GRB afterglows have very low column densities. A large sample of high-resolution spectra of GRB afterglows could provide statistical information about the distribution of the gas in high-redshift star-forming regions, in addition to the evolution of the metallicity and dust and H$_2$ contents of GRB host galaxies. Such a sample can be created in the years to come thanks to rapid and accurate GRB localizations from future satellite missions such as Swift\\footnote{see http://swift.gsfc.nasa.gov/} and EXIST\\footnote{see http://exist.gsfc.nasa.gov/}." }, "0403/astro-ph0403205_arXiv.txt": { "abstract": "We present the $UBVRI$ CCD photometry in the region of the open cluster NGC 2421. Radius of the cluster is determined as $\\sim$ 3$^\\prime$.0 using stellar density profile. Our Study indicates that metallicity of the cluster is $Z \\sim$ 0.004. The reddening $E(B-V) = 0.42\\pm$0.05 mag is determined using two colour $(U-B)$ versus $(B-V)$ diagram. By combining the 2MASS $JHK$ data with the optical data we determined $E(J-K) = 0.20\\pm$0.20 mag and $E(V-K) = 1.15\\pm$0.20 mag for this cluster. Colour-excess diagrams show normal interstellar extinction law in the direction of the cluster. We determined the distance of the cluster as 2.2$\\pm$0.2 Kpc by comparing the ZAMS with the intrinsic CM diagram of the cluster. The age of the cluster has been estimated as 80$\\pm$20 Myr using the stellar isochrones of metallicity $Z = 0.004$. The mass function slope $x = 1.2\\pm0.3$ has been derived by applying the corrections of field stars contamination and data incompleteness. Our analysis indicate that the cluster NGC 2421 is dynamically relaxed. ", "introduction": "The investigation of young open star clusters provides us a powerful tool to understand the structure and history of star formation in our Galaxy. In order to fully exploit the information provided by open clusters we must know their accurate ages, distances, reddenings, metal abundances and stellar contents. For this, multicolour CCD photometric observations have proved to be very useful. With the development of more accurate stellar models it has been possible to provide a better estimate of the cluster ages simply by comparing theoretical isochrones with the observed CCD colour-magnitude (CM) diagrams. So, In this paper we have considered an open cluster NGC 2421 with the aim of presenting new accurate CCD photometry. From this photometry we select photometric members and derive several fundamental parameters, such as distance, interstellar reddening, metallicity and age as well as luminosity and mass function. The young open cluster NGC 2421 = C0734$-$205 ($\\alpha_{2000} = 07^{h}36^{m}16^{s}$, $\\delta_{2000}=-20^{d}36^{\\prime}44^{\\prime\\prime}$; $l = 236^{\\circ}.24$, $b = 0^{\\circ}.08)$ is classified as a Trumpler class I2m by Ruprecht (1966). This cluster was first studied by Moffat \\& Vogt (1975) photoelectrically and derived a distance of about 1.87 Kpc having $E(B-V)=$ 0.47$\\pm$0.05 mag and age less than 10$^{7}$ years. Ramsay \\& Pollacco (1992) also studied this cluster using CCD photometry and found a colour excess $E(B-V)=$ 0.49$\\pm$0.03 mag but a distance of 2.75 Kpc. To our knowledge no other studies have been carried out for the cluster NGC 2421 so far. The layout of the paper is as follows. In Sec. 2 we briefly describe the observations and data reduction strategies as well as comparison with the previous photometry. Sec. 3 is devoted on the detailed analysis of the present photometric data for the determination of cluster parameter. Finally, Sec. 4 summarizes the main results of the paper. \\begin{figure*} \\centering \\hspace{0.7cm}\\psfig{figure=ME150fig1.ps,width=10cm,height=10cm} \\caption{Finding chart of the stars in the cluster NGC 2421. The (X, Y) coordinates are in pixel units corresponding to 0$^{\\prime\\prime}$.72 on the sky. Direction is indicated in the map. Filled circles of different sizes represent brightness of the stars. Smallest size denotes stars of $V$$\\sim$20 mag. Open circle having centre at 'C' in the chart represent the cluster size.} \\end{figure*} ", "conclusions": "We have investigated the area of open cluster NGC 2421 using $UBVRI$ CCD and 2MASS $JHK_s$ data. The main results of our analysis are the following. \\begin{enumerate} \\item The radius of the cluster is determined as 3$^\\prime$.0 which corresponds to 1.9 pc at the distance of the cluster. \\item We estimated the abundance of the cluster stars as $Z =$ 0.004 using excess in $(U-B)$. The $(U-B)$ versus $(B-V)$ colour-colour diagram yields $E(B-V) = 0.42\\pm$0.05. The analysis of $JHK$ data in combination with the optical data provide $E(J-K) = 0.20\\pm$0.20 mag and $E(V-K) = 1.15\\pm$0.20 mag. Our analysis shows that interstellar extinction law is normal towards the cluster direction. No stars found which are having near-IR fluxes due to the presence of circumstellar material around them. \\item A ZAMS fitting procedure gives a distance of 2.2$\\pm$0.2 Kpc for this cluster which is also supported by the value of 2.3$\\pm$0.3 Kpc determined by us using the optical and near-IR data. An age of 80$\\pm$20 Myr is determined using the isochrones of $Z =$ 0.004 given by Bertelli et al. (1994). \\item The mass function slope $x = 1.2\\pm0.3$ is derived by considering the corrections of field star contamination and data incompleteness. Our analysis indicate that the cluster NGC 2421 is dynamically relaxed and one plausible reason of this relaxation may be the dynamical evolution of the cluster. \\end{enumerate}" }, "0403/astro-ph0403033_arXiv.txt": { "abstract": "{ We report the results of a non-LTE Fe, O and Mg abundance analysis of the carbon-nitrogen-rich ultra-metal-poor giants CS\\,29498--043 and CS\\,22949--037. The abundance of oxygen has been derived from measurements of the oxygen triplet at 7771--5 \\AA\\ in high resolution spectra obtained with KeckI/HIRES and the forbidden line [O\\,{\\sc i}] 6300 \\AA\\ detected in the TNG/SARG spectra of CS\\,29498-043. Detailed non-LTE analysis of Fe lines has provided reliable stellar parameters which, however, do not resolve the oxygen abundance conflict as derived from the O\\,{\\sc i} 7771-5 \\AA\\ triplet and the [O\\,{\\sc i}] 6300 \\AA\\ forbidden lines. We obtained the following oxygen abundance: for CS\\,22949--037 [O/Fe] =3.13, 1.95; and for CS\\,29498--043; [O/Fe]=3.02, 2.49, based on the O\\,{\\sc i} 7771--5 \\AA\\ triplet and the [O\\,{\\sc i}] 6300 \\AA\\ forbidden line, respectively. A similar conflict appears to exist between the forbidden resonance line Mg\\,{\\sc i} 4571 \\AA\\ and several subordinate lines, such as Mg\\,{\\sc i} 5172 and 5183 \\AA. Our analysis demonstrates the failure of standard plane--parallel atmosphere models to describe the physical conditions in the line-forming regions of these ultra-metal-poor giants. ", "introduction": "Surveys of metal-poor stars (Beers et al. 1992) and their abundance studies (Ryan, Norris \\& Beers 1996; Norris, Ryan \\& Beers\\,1999) are aimed at investigating the chemical evolution of the Galaxy and the nucleosynthetic yields of supernovae. Despite the numerous studies in this field, the current situation with abundance trends of various $\\alpha$-elements is very confusing. Observations of Stephens \\& Boesgaard (2002) demonstrate that the [$\\alpha$/Fe] ratios for Ca, Si, Ti and Mg do not show a flat {\\itshape plateau} at [Fe/H] $< -1$, as was claimed in previous studies. The results obtained by Idiart \\& Thevenin\\,(1999) and Stephens \\& Boesgaard (2002) cannot be called {\\it consistent} with the analysis presented by Carretta et al.\\,(2002) and many others (see McWilliam 1997). The situation with oxygen is far from being resolved (Israelian, Garc\\'\\i a L\\'opez \\& Rebolo \\,1998; Israelian et al.\\,2001; Nissen et al.\\ 2002; Takeda 2003; Fulbright \\& Johnson\\ 2003), while a new debate over the sulfur abundance in metal-poor stars has already emerged (Israelian \\& Rebolo 2001; Takada-Hidai et al.\\ 2002; Nissen et al.\\ 2003). These and many other studies clearly show that [$\\alpha$/Fe] $>$ 0 for the great majority of metal-poor stars in the Galaxy. However, it is hard to speak of any trend when the abundance ratios in many stars computed by different authors disagree by more than 0.3--0.4 dex. There are many aspects to this serious problem, and we shall not discuss them further here. Abundance analysis of ultra-metal-poor stars with [Fe/H] $< -$3 gives rise to even more enigmas into this field. It is well known that the chemical composition of the atmospheres of halo dwarfs is not altered by any internal mixing and therefore provides a good opportunity to constrain Galactic chemical evolution models. Unfortunately, most of the known ultra-metal-poor stars are not dwarfs but giants (Beers et al. 1992), which pose two serious problems. First, the atmospheric parameters of giants are more uncertain and second, their surfaces can be polluted by enriched material that has been either dredged from the stellar interior or transferred from a companion star. Oxygen is a key element in this scheme as it can help to distinguish between pristine and pollution origins of other elements and also show which of the aforementioned processes was dominant. There have been intensive investigations of the oxygen abundances in halo stars over the last five years. Abundances derived from near-UV OH lines in metal-poor dwarf stars (Israelian et al.\\,1998; Boesgaard et al. 1999; Israelian et al.\\,2001) show that the [O/Fe] ([O/Fe] = log(O/Fe)$_\\star$--log (O/Fe)$_\\odot$) ratio increases from 0 to 1 between [Fe/H] = 0 and $-3$. The abundances derived from low-excitation OH lines agreed well with those derived from high-excitation lines of the O\\,{\\sc i} triplet at 7771--5 \\AA\\ (Israelian et al. 1998, 2001; Boesgaard et al. 1999 ; Nissen et al.\\,2002). It seems that even the [O\\,{\\sc i}] forbidden line at 6300 \\AA\\ supports the ``quasi-linear'' trend of [O/Fe] (Nissen et al.\\,2002) when standard 1D atmospheric models are employed. While some authors claim a good agreement between the forbidden line [O\\,{\\sc i}] 6300 \\AA\\ and the near-IR triplet (Mishenina et al. 2000; Nissen et al.\\,2002), others suggest the opposite (Carretta, Gratton, \\& Sneden 2000). In a recent study Takeda (2003) found that the disagreement between the triplet and the forbidden line tends to be larger for cool giants. Fulbright \\& Johnson (2003) support this conclusion in a detailed study of 55 subgiants and giants. These authors conclude that it is impossible to resolve the disagreement in the two indicators without adopting an ad hoc temperature scale that is incompatible with standard temperature scales such as IRFM and H$\\alpha$. In fact, the [O/Fe] trend obtained based on the ad hoc scale (see Fig.\\ 13 of Fulbright \\& Johnson 2003) is identical to the trends presented by Israelian et al. (2001) and Nissen et al. (2002). There is no {plateau} at [O/Fe] = 0.5. Despite considerable observational effort the trend of the [O/Fe] ratio in the halo is still unclear. However, the latest studies (Israelian et al. 2001; Nissen et al. 2002; Takeda 2003; Fulbright \\& Johnson 2003) suggest that dwarfs provide more reliable and consistent abundances than giants. It is not clear how the 3D effects will resolve this conflict since the latter do not predict an agreement between different oxygen abundance indicators in dwarfs (Asplund \\& Garc\\'\\i a P\\'erez 2002). McWilliam et al.\\,(1995) were the first to carry out a detailed spectroscopic analysis of CS\\,22949-037 and to confirm that the star is very metal-poor with an $\\alpha$-element excess. Furthermore, Depagne et al.\\ (2002) performed a more detailed investigation of this object and found a large excess of oxygen ([O/Fe] = 2.0) and sodium ([Na/Fe] = 2.1). Zero-heavy-element supernovae models with fall back have been invoked in order to interpret the elemental abundance ratios in this star. Aoki et al.\\, (2002) have presented a detailed analysis of another ultra-metal-poor giant CS\\,29498--043 with a very high abundance excess of [Mg/Fe] = 1.81. Both, CS\\,22949--037 and CS\\,29498--043 exhibit a large overabundance of N and C but show no significant enhancement of neutron-capture elements. It is possible that the surfaces of these stars have been polluted by enriched material, either dredged from the star's inner core or transferred from a companion star. The abundances of $\\alpha$-elements could be used to discriminate in favour of one of these hypotheses or to confirm a pristine origin. The detailed comparison of elemental abundances may provide important constraints on the properties of the first supernova progenitors. In this article we present observations of the oxygen triplet in CS\\,22949--037 and CS\\,29498-043, as well as the detection of the forbidden line [O\\,{\\sc i}] 6300 \\AA\\ in CS\\,29498--043. The stellar parameters and the abundances of iron, oxygen and magnesium were derived in non-LTE. We report a significant discrepancy between the abundances derived from the oxygen triplet and the forbidden line. The conflict cannot be resolved under any circumstances, at least for CS\\,22949-037. A similar conflict was found for Mg. We question the validity of standard plane--parallel models of atmospheres employed in the present analysis. ", "conclusions": "Observations with Keck I/HIRES have revealed strong lines of the oxygen near-IR triplet in the spectra of the ultra-metal-poor giants CS\\,22949--037 and CS\\,29498--043. The forbidden line of oxygen with EW = 60$\\pm$10m\\AA\\ was observed with TNG/SARG. A detailed non-LTE analysis of Fe has been carried out and a new set of the atmospheric parameters have been obtained. Our analysis suggests that the gravities of metal-poor giants derived from the LTE Fe analysis are strongly underestimated because of the neglect of non-LTE effects. The oxygen abundance in CS\\,22949--037 and CS29498--043 derived from the triplet and the forbidden line differ by 1.18 and 0.53 dex, respectively. This disagreement cannot be explained by a non-LTE effects, quality of the data and/or uncertainties in stellar parameters. Other mechanisms must be invoked in order to explain this puzzle. A similar discrepancy was found for Mg when comparing the abundances obtained from the resonance line 4571 \\AA\\ and several strong subordinate lines. Based on the present analysis we propose that the Kurucz (1992) models are not reliable for these ultra-metal-poor giants." }, "0403/astro-ph0403519_arXiv.txt": { "abstract": "We present VLT-UVES echelle spectra of the \\Lya~emission and absorption in five radio galaxies at redshifts z=2.55--4.1. Together with data from our pilot study, we have a sample of 7 such systems with radio source sizes $\\sim 1-90$\\kpc~with which to address the origin of the absorbing gas. Echelle resolution again reveals that some systems with $N_{\\rm{HI}}>10^{18}$\\psqcm~in lower resolution data in fact consist of several weaker absorbers with $N_{\\rm{HI}}<10^{15}$\\psqcm. We identify two groups of HI absorbers: strong absorbers with $N_{\\rm{HI}} \\simeq 10^{18}-10^{20}$\\psqcm~and weaker systems with $N_{\\rm{HI}} \\simeq 10^{13}-10^{15}$\\psqcm. There are none at intermediate $N_{\\rm{HI}}$. The strong absorbers may be a by-product of massive galaxy formation or could instead represent material cooling behind the expanding bow-shock of the radio jet, as simulated by Krause. New observations are required to discriminate between these possibilities. We argue that the weaker absorbers with $N_{\\rm{HI}} \\simeq 10^{13}-10^{15}$\\psqcm~are part of the \\Lya~forest, as their rate of incidence is within a factor of 2--4 of that in the IGM at large. Such column densities are consistent with models of a multi-phase proto-intracluster medium at $z>2$. ", "introduction": "High-redshift ($z>2$) radio galaxies (HzRGs) are among our best probes of massive galaxy formation. The K-z relation demonstrates that their supermassive black holes ($10^{9}-10^{10}$\\Msun) reside in the most massive elliptical galaxies at their epoch (e.g. Jarvis et al.~2001a; Willott et al.~2003). Measurements of galaxy clustering and of galaxy over-densities around individual sources, imply that they inhabit proto-cluster environments which will evolve into rich clusters at $z=0$ (for a review of radio source clustering and proto-clusters, see R\\\"{o}ttgering et al.~2003). Deep within these gravitational potential wells and surrounding the HzRGs to radii of tens to several hundred \\kpc, are luminous ($>10^{44}$\\ergps~in \\Lya) emission line nebulae. Their properties bear witness to the complex gaseous processes in nascent massive galaxies, viz. gravitational collapse and shock heating, cooling, bursts of star formation and feedback in the form of superwinds; the latter can inject substantial quantities of energy and metals into the gas, enriching it to solar or super-solar metallicity out to tens of \\kpc~(e.g. Villar-Martin et al.~2001). The presence of a powerful radio source has a profound impact on such an environment: within the dense interstellar medium of the galaxy, the relativistic radio jet drives shocks into the clumpy gas, leading to extreme kinematics ($FWHM > 1000$\\kmps; Villar-Martin, Binette and Fosbury~1999), additional star formation and the emission line signatures of shock ionisation (Best et al.~2000; De Breuck et al.~2000). At radii beyond the radio source, the gas is kinematically quiescent (reflecting purely gravitational motions) and the continuum from the quasar nucleus is the main source of ionisation (Villar-Martin et al.~2003). New insight into the gaseous haloes of HzRGs came with the discovery by R\\\"{o}ttgering et al.~(1995) and van Ojik et al.~(1997) that the majority of small radio sources (those with projected linear size below 50\\kpc) exhibit spatially resolved absorption in their \\Lya~profiles, with HI column densities of $10^{18}-10^{19.5}$\\psqcm. For the $z=2.92$ HzRG 0943--242, Binette et al.~(2000) discovered associated CIV$\\lambda\\lambda$1548,1551~absorption at the same redshift as the main \\Lya~system. Through photoionisation modelling they argued that the absorbing gas is of lower metallicity ($Z \\sim 0.01Z_{\\rm{\\odot}}$) and is situated beyond the emission line gas, outside the high pressure radio source cocoon. This material, they claimed, is a relic reservoir of low metallicity, low density ($n \\sim 10^{-2.5}$\\pcm) gas. In larger radio sources, the radio source cocoon plausibly engulfs and pressurises the outer halo entirely, causing it to be seen in emission rather than absorption, and thereby accounting for the anti-correlation between radio source size and the presence of HI absorption found by van Ojik et al.~(see also Jarvis et al.~2001b for the HI absorption:radio source size correlation). With the Ultraviolet-Visible Echelle Spectrograph (UVES) at the VLT, we recently examined the absorbers in two HzRGs with over ten times the spectral resolution of R\\\"{o}ttgering et al. and van Ojik et al. Our targets were the aforementioned 0943--242, and 0200+015 ($z=2.23$) and the results were presented by Jarvis et al.~(2003) and Wilman et al.~(2003). In 0943--242, the absorbers exhibit no additional structure -- the main absorber is still consistent with an HI column density of $10^{19}$\\psqcm, in line with the Binette et al.~(2000) model. In contrast, a very different view of 0200+015 emerges: the single absorber with HI column density $\\simeq 10^{19}$\\psqcm seen at low resolution now splits into two $\\sim 4 \\times 10^{14}$\\psqcm~systems; these extend by more than 15\\kpc~to obscure addtional \\Lya~emission coincident with a radio lobe and highly fragmented absorption is also seen on the red wing of the emission line at this position. The detection of one of the sub-systems in CIV suggests that the absorbing gas has undergone substantial metal enrichment, to perhaps as high as $10Z_{\\rm{\\odot}}$. Based on the smaller radio source size in 0943--242 (26\\kpc~versus 43\\kpc~for 0200+015), we conjectured that the radio source age (as inferred from its linear size) is the parameter controlling the evolution of: (i) the structure/kinematics of the absorbing halo, through interaction and shredding of the initially quiescent shells; (ii) its metallicity, through enrichment by starburst superwind triggered concurrently with the nuclear radio source. If the specific effects of the radio source can be understood, high resolution spectroscopy of such \\Lya~haloes provides an opportunity to probe the \\Lya~forest around high-redshift galaxies and hence to test theories for how such systems form out of the intergalactic medium (IGM). For quasars, such attempts are completely hampered by their strong UV continua, which through photoionisation locally reduce the number of \\Lya~forest clouds below the IGM average (the so-called proximity effect). Lacking strong line of sight continua, radio galaxies offer a more direct probe of the distribution of HI in such dense environments. Due to the nature of hierachical clustering, the opacity of the \\Lya~forest is expected to increase in the vicinity of massive objects, and observations of background quasars do indeed show excess HI absorption at radii of 1--5$h^{-1}$\\Mpc~from $z \\sim 3$ Lyman-break galaxies (LBGs) (Adelberger et al.~2003). However, within $0.5h^{-1}$\\Mpc~there appears to be a deficit of HI, which they attribute to the influence of `superwind' outflows from the LBGs. From a theoretical stand-point it is not clear that superwinds alone can account for this observation, and additional effects such as local photoionization (Bruscoli et al.~2003) and the `filling in' of absorption with \\Lya~emission from the LBG itself (Kollmeier et al.~2003) may be needed. In order to address the origin and evolution of the absorbing shells and their connection with the \\Lya~forest, we need to expand our sample of HzRGs with high resolution spectroscopy. For this reason, we have recently acquired UVES spectra for a further 5 HzRGs, yielding a sample of 7 HzRGs with radio source sizes in the range $\\sim 1-90 $\\kpc~(projected). Here we present the analysis and interpretation of the new dataset. The paper is structured as follows: in section 2, we describe the target selection, observations and data reduction, and in section 3 present the \\Lya~profiles and absorption model fits on an object-by-object basis. In section 4 we compile numerous statistics for the absorption line ensemble and compare them with the properties of HI absorption in the IGM. In section 5, we interpret these results in the light of simulations of: (i) expanding radio sources in dense environments; (ii) the \\Lya~forest around high-redshift galaxies. Section 6 contains our conclusions. Throughout the paper we assume a cosmology with $H_{\\rm{0}}=70$\\kmpspMpc, $\\Omega_{\\rm{M}}=0.3$ and $\\Omega_{\\rm{\\Lambda}}=0.7$. ", "conclusions": "We have built on our successful pilot study and increased the number of high-redshift radio galaxies with UVES echelle spectroscopy of \\Lya~from 2 to 7. With this enlarged ensemble of absorption systems we can begin to make some inferences about their origin. The new spectra demonstrate once again that HI absorption is more complex when seen at this much higher resolution: several absorbers identified in low-resolution spectra with $N_{\\rm{HI}}>10^{18}$\\psqcm~now fragment into a number of much weaker systems with $N_{\\rm{HI}}<10^{15}$\\psqcm. There appears to be a gap in the $N_{\\rm{HI}}$ distribution, as we identify strong absorbers with $N_{\\rm{HI}} \\simeq 10^{18}-10^{20}$\\psqcm~and weaker systems with $N_{\\rm{HI}} \\simeq 10^{13}-10^{15}$\\psqcm, but none at intermediate $N_{\\rm{HI}}$. We discussed the origin of the strong absorption systems with $N_{\\rm{HI}}>10^{18}$\\psqcm~in the framework put forward by Krause~(2002), based on radio source physics. A highly supersonic jet expanding into a dense proto-galactic medium will be surrounded by a quasi-spherical bow shock. The simulations by Krause~(2002) show that the cooling of post-shock material will lead to neutral column densities comparable to those observed, and that due to instabilities the shells will fragment and offer much less absorption towards larger radio sources. The model provides a very natural explanation for the observation that the shells have high covering factors and surround the entire galaxy. Appealing though it is, it would be premature to enshrine it as the canonical model for the strong absorbers in HzRGs. Several new challenging observations are needed before we can determine whether such strong absorbers really are produced by the radio source bow shock or whether they are instead a by-product of massive galaxy formation. Firstly, in the HzRGs themselves deep, high resolution spectra should be obtained of the faint kinematically-quiescent \\Lya~haloes, identified by Villar-Martin et al.~(2003). Secondly, such spectroscopy should also be obtained of a carefully selected sample of comparably massive high-redshift galaxies without radio-loud AGN, comprising SCUBA galaxies and narrow-line X-ray type II quasars. We cannot exclude the possibility that some of the lower column density systems with $N_{\\rm{HI}}=10^{13}-10^{15}$\\psqcm~form by fragmentation of larger shells until the simulations of the bow-shock absorbers are extended into this regime. The rate of incidence of such absorption in our spectra and their distribution on the $b$--$N_{\\rm{HI}}$ plane suggests, however, that these systems could be part of the IGM \\Lya~forest population. Simulations predict that the opacity of the forest should increase within a few Mpc~of a massive galaxy at these epochs. Our results show an enhancement (by factors 2--4) in the number of \\Lya~forest systems (relative to the IGM) at $z=2-3$ and a deficit by similar factors in the two $z>3.5$ systems, although on the grounds of small number statistics we do not claim that this implies evolution over this redshift interval. For a proper comparison, independent of any Voigt fitting uncertainties and systematics, new simulations of the \\Lya~forest should be performed. Unlike those in the literature which predict the HI opacity at the galaxy redshift as a function of transverse distance from it, the new simulations should yield the \\Lya~forest along the line of sight to a massive object in redshift space, as a function of galaxy mass and epoch. In lieu of this, we note that the observed column densities are consistent with simple estimates for a multiphase proto-intracluster medium at $z>2$." }, "0403/astro-ph0403343_arXiv.txt": { "abstract": "We present \\chandra\\ X-ray observations of the young supernova remnant (SNR) \\e05 in the Large Magellanic Cloud (LMC), believed to be the product of a Type Ia supernova (SN Ia). The remnant is very round in shape, with a distinct clumpy shell-like structure that extends to an average radius of $14.8\\arcsec$ (3.6 pc) in the X-ray band. Our \\chandra\\ data reveal the remnant to be rich in silicon, sulfur, and iron. The yields of our fits to the global spectrum confirm that \\e05 is the remnant of an SN Ia and show a clear preference for delayed detonation explosion models for SNe Ia. The \\chandra\\ spectra extracted from radial rings are in general quite similar; the most significant variation with radius is a drop in the equivalent widths of the strong emission lines right at the edge of the remnant. We study the spectrum of the single brightest isolated knot in the remnant and find that it is enhanced in iron by a factor of roughly two relative to the global remnant abundances. This feature, along with similar knots seen in Tycho's SNR, argues for the presence of modest small-scale composition inhomogeneities in SNe Ia. The presence of both Si and Fe, with abundance ratios that vary from knot to knot, indicates that these came from the transition region between the Si- and Fe-rich zones in the exploded star, possibly as a result of energy input to the ejecta at late times due to the radioactive decay of $^{56}$Ni and $^{56}$Co. Two cases for the continuum emission from the global spectrum were modeled: one where the continuum is dominated by hydrogen thermal bremsstrahlung radiation; another where the continuum arises from non-thermal synchrotron radiation. The former case requires a relatively large value for the ambient density ($\\sim$1 cm$^{-3}$). Another estimate of the ambient density comes from using the shell structure of the remnant in the context of dynamical models. This requires a much lower value for the density ($<$0.05 cm$^{-3}$) which is more consistent with other evidence known about \\e05. We therefore conclude that the bulk of the continuum emission from \\e05 has a non-thermal origin. ", "introduction": "Type Ia supernovae (SN Ia) are characterized by the absence of Balmer lines and the presence of strong silicon lines in their optical spectra \\citep{hillnie}. They yield mostly intermediate mass elements from Si to Ca and are the main sources of iron production in galaxies. Each SN Ia produces $\\sim$0.8 ${\\rm M{_\\sun}}$ of Fe \\citep{iwam}, making them important for understanding the chemical evolution of galaxies. The peak magnitude and shape of their light curves are fairly homogeneous, which makes them useful as distance indicators and tools for cosmology \\citep{rpk,riess,perl}. For these reasons, SN Ia's are the objects of much study. Still, there is much that remains uncertain about them. The progenitors of SN Ia's are believed to be degenerate C+O white dwarfs. Although models for sub-Chandrasekhar mass explosions exist, the favored explanation is that the degenerate object somehow exceeds the Chandrasekhar limit before exploding. Two possibilities include a single degenerate scenario, in which the white dwarf accretes mass from a companion star, and a double degenerate scenario, in which a pair of white dwarfs coalesce \\citep{branch}. It is still unclear whether one, both, or none of these offers the correct explanation. In addition, the mechanism of the explosion, i.e., where the ignition begins and how it propagates outward, continues to elude us \\citep{iwam}. Two cases studied in detail include a fast deflagration and a delayed detonation. In the former, a deflagration begins close to the center of the progenitor star and propagates subsonically. The delayed detonation case involves a deflagration initially, which is transformed at some density to a detonation as the star expands. Models for these two cases are parameterized by the speed of the burning front and the location of the ignition in density space; we compare our results with those of \\citet{iwam}. By studying the remnants of SN Ia explosions, we can hope to gather clues to some of these puzzles. The supernova remnant (SNR) \\e05 in the Large Magellanic Cloud (LMC) is believed to be the remnant of an SN Ia \\citep{tuohy, hughes95}. It was discovered in an X-ray survey of the LMC with the {\\it Einstein X-ray Observatory} \\citep{long}. However, much of the previous study on this remnant was done via optical spectroscopy. In H$\\alpha$, the remnant appears as a very circular shell \\citep{tuohy}. \\e05 shows no detectable emission in [\\ion{O}{3}] and [\\ion{S}{2}], indicating it is a Balmer-dominated remnant \\citep{tuohy, smith91,smith94}. Remnants such as this generally have two components to the H$\\alpha$ emission line: a narrow component produced by neutral hydrogen being excited by the shock, and a broad component produced by charge exchange with fast protons post-shock. In \\e05, Smith et al. (1994) found the narrow components to be $\\sim$25--31 km/s, higher than expected for neutral hydrogen and a possible indication of the presence of a cosmic ray precursor. This SNR is also intriguing in that its broad component is believed to be so broad that it has escaped detection, pointing to very fast shocks ($\\geq$2000 km/s) and a young age ($\\leq$1000 yr) (Smith et al. 1991; Smith et al. 1994). \\citet{tuohy} found that \\e05 is expanding into a low density medium, with $n_{\\rm H}\\lesssim0.02$ cm$^{-3}$. A previous X-ray study of this SNR with {\\it ASCA} revealed it to be strong in Si and Fe L emission, with little O, Ne, and Mg emission \\citep{hughes95}. This remnant has also been observed in the radio, showing a spectral index of $\\alpha=0.46$ \\citep{mathewson} and a flux density of 0.066 Jy at 1 GHz \\citep{henrey}. ", "conclusions": "Our analyses of \\e05 point us toward several conclusions which give us insight into various aspects of SNRs. We find that the abundances obtained from our model fits, when compared with theoretical models of supernovae, confirm that \\e05 is the remnant of an SN Ia. In addition, the abundance fits are able to give us information as to which theoretical model is more appropriate. The global spectrum of \\e05 as well as that of a small knot of material give us important clues as to the properties of iron in the remnant. Our deprojection and image fits provide us with a shell structure to this SNR, which we use to derive information about the dynamics. This in turn allows us to draw the conclusion that the bulk of the continuum emission in this remnant must be non-thermal in nature. \\subsection{SN Ia Yields} In Figure 8 we plot our abundances, normalized to Si = 1, for Cases H (squares) and S (x's). These yields are also compared to those determined by two models of SN Ia, the W7 model (fast deflagration, stars) and the WDD3 model (delayed detonation, circles) of \\citet{iwam}. The fast deflagration involves a deflagration wave that propagates outward subsonically, and the delayed detonation scenario begins as a deflagration that then transitions to a detonation at some density. The other models presented by \\citet{iwam} are effectively represented here. Their W70 model is similar to W7, and the other DD models are similar to WDD3; therefore we did not plot these. It is quite clear that the low-Z elements of O, Ne and Mg are 1--2 orders of magnitude less abundant than the higher-Z elements of Si - Ca. This is a signature of SN Ia's \\citep{thiele} and confirms such an origin for this remnant. Our yields for O, Ne, and Mg agree better with the WDD3 (delayed detonation) model than the W7 (fast deflagration) model. Even if all of the Si has not been shocked, the low O/Si ratio would only be exacerbated and the W7 model would still fail to fit the data. \\citet{badenes} have compared various explosion models with the spectrum of Tycho's SNR. They are able to show that the pure detonation and pure deflagration models do not satisfactorily reproduce important features in Tycho's spectrum. Their results point to a delayed detonation explosion as the best model. In their Figure 4, they plot the integrated EM of each species in the remnant as a function of time. In this context, we looked at another Type Ia remnant, DEM L71. This SNR is 4400 years old \\citep{ghav} and rich in Si, S, and Fe \\citep{hughl71}. Shocks in older remnants probe different parts of the SN ejecta. In comparison to their Figure 4, we find that at the age of DEM L71 the models which show the same dominant elements as seen in this SNR are also delayed detonation models. This suggests that the delayed detonation scenarios for SN Ia explosions may be more appropriate. In addition, all three remnants indicate some preference for higher transition densities among the delayed detonation models. All of these results reveal the power of remnant studies of Ia's to elucidate explosion mechanism physics. The iron yields we observe are low compared to the models of \\citet{iwam}. Our model does underestimate the amount of iron present in the remnant. However, it is extremely unlikely that uncertainties in Fe L-shell atomic physics could account for the difference of a factor of $\\sim$30 between the models and our yield. We interpret this discrepancy to be due to the fact that the reverse shock has not propagated far enough into the remnant to shock all of the iron. This is not unprecedented; there is evidence for unshocked iron ejecta in another young Type Ia, SN1006, where \\ion{Fe}{2} absorption lines have been detected in the center of the remnant \\citep{hamilton, wu}. This is also a feature of theoretical models \\citep{badenes}. One puzzle we encountered was that our fits required iron to be in a different thermodynamic state than the other elements, with a much higher temperature and slightly higher ionization timescale. We attempted a fit with a component of iron at the same state as the other elements and found we could allow a maximum abundance of 0.85 for both Cases H and S, which is 0.05 and 0.07, respectively, relative to Si. However, such a fit did not reproduce the Fe K line and quite clearly underestimated the flux from 0.9--1.4 keV. The fact that iron appears to be in a different thermodynamic state implies that it may have come from a different portion of the progenitor than the other elements. \\citet{hwang} found a similar situation in the case of Tycho's SNR, where the Fe K emission required a higher temperature, but a lower ionization timescale than the other elements. Intuitively, this might be what one would expect if there were a temperature gradient and the ejecta were stratified, with Fe interior to the other elements. We may be seeing an indication of the ejecta density profile here. For example, the constant density ejecta profile of \\citet{dc} predicts an increasing temperature gradient behind the contact discontinuity. However, it must be emphasized that X-ray studies probe the electron temperature, while the main thermal energy rests in the ions. The extent of the equilibration between these two components is still under investigation (e.g., Rakowski, Ghavamian, \\& Hughes 2003). In addition, it is not simple to relate the values fit for in our models to the {\\em actual} temperature and timescales of the ejecta because of astrophysical complications. For example, our fits assume a single, constant temperature for each species everywhere in the remnant. \\subsection{Iron Enhanced Knot} We have determined that the knot in the northern interior of \\e05 is enhanced in iron (see Figure 9). As seen in projection, this knot is contained mostly in the second northeast half-ring. We used the spectrum of that particular ring as a template for the knot's spectrum. If we simply renormalize the spectrum, the fit is poor ($\\chi^2/d.o.f. = 59.0/32$). However, if we allow the iron abundance to be free, we find a good fit at about twice the iron abundance of the half-ring ($\\chi^2/d.o.f. = 32.6/32$). A fit of pure iron also gave a poor match to the data, indicating that although this knot is enhanced in iron, it is not pure iron ejecta. We looked at this particular knot simply because it was the brightest and most isolated. However, given that our image fits (\\S 3.2) show the presence of significant clumpiness in this SNR, it is possible that other knots show variations in the Si/Fe abundance ratio, although we are unable to definitively measure this. Knots with different X-ray spectra have been detected in Tycho, the so-called ``typical'' Type Ia SNR, on the southeastern edge of the remnant \\citep{van}. Recent work has confirmed that the spectral differences are a result of differences in Si/S and Fe abundances in these knots \\citep{dec}. In neither Tycho nor \\e05 do we see knots of {\\em pure} iron or silicon. Models of SN Ia explosions (i.e., Iwamoto et al. 1999) predict that the ejecta are highly stratified, with an outer O-rich layer, then a Si-rich region, followed by an inner Fe-rich layer. However, the boundaries between these regions are not sharp; the abundances vary smoothly from one zone to the next. The differing ratios of Si/Fe observed in the remnants' knots suggest to us that they originated in the transition region between the iron- and silicon-rich zones of the ejecta. Clumping in SN Ia's has been investigated by \\citet{wc}. They propose that the nickel bubble effect, in which the radioactive nickel expands and compresses the shell of material around it, may be responsible for producing knots. The shell surrounding the nickel would be the Si/Fe transition zone. Further investigation into possible origins of knots in this region would be worthwhile. \\subsection{Shell Structure} Our deprojection analysis clearly indicates the presence of a true shell structure for this SNR (see Figure 7). The deprojection was done for O, Si, S, Fe, and H (the latter only for Case H). In both cases, and for both hemispheres, we found that the metals all follow this structure, with essentially no material in the interior of the remnant (see below for a discussion of the southwest hemisphere). The metals extend over the two outermost shells in the deprojection with a peak density in the shell at $R \\sim 3.2$ pc and a slightly lower density in the outer shell at $R \\sim 3.6$ pc. Hydrogen (Case H) shows a slightly different structure: there is almost no H in the $R \\sim 3.2$ pc shell or further in the interior; rather the continuum comes from the outermost shell. The continuum emission in Case S, when deprojected, also comes only from the outermost shell. In each case, therefore, this is clear evidence for a blast-wave component to the remnant. We note however that the blast wave component appears to be contaminated somewhat by ejecta, presumably by fragments that preceed the main shell. In the southwest hemisphere, the structure is more complicated. There is a shell at a radius of $\\sim$ 3.2 pc that is the counterpart to the northeast shell. In addition, there appears to be another spectral component peaking at a smaller radius, with an apparent gap between the two shells. We do not believe this is an actual shell, but rather a projection effect due to the clumpy region on the southwest side. Given simply a shell of material, one would expect to see a radial profile like that of the northeast hemisphere. Now if one adds a clump interior to the shell, the radial profile would include a ``bump'' at some inner radius, as seen in the southwest hemisphere. The radius at which this bump appears gives no indication of the true position of the clump, which could be almost anywhere along the line-of-sight in the remnant. Our simple deprojection analysis ignores this fact, and assumes that the emission from the clump is spread over an inner shell. Therefore, the density derived for these inner portions is clearly an overestimate for the interior of the remnant. On the other hand, it may be an underestimate for the true density of the clump itself. We believe that the clumpy region in the southwest of \\e05 is a result of enhanced density in or a deeper penetration of the reverse shock into a portion of the ejecta shell. This could be caused by a small region of enhanced ambient density or by intrinsic asymmetry in the explosion process itself. Because our deprojection analysis revealed the shell-like structure of this remnant, we can be confident that the image fits, which assumed such a structure, produced meaningful results. The fit results give us a constraint on the evolutionary state of the remnant if we identify the outer edge of the X-ray emitting shell with the location of the blast wave, and the inner edge with the location of the reverse shock. In the context of self-similar models for the evolution of young SNRs (e.g., Truelove \\& McKee 1999), the observed ratio of these radii (1.185--1.192, including errors), corresponds to a particular point (or range of points) on the scaled radius vs.\\ time curve. These curves depend on the assumed radial density distribution of the ejecta; we consider the cases of constant and exponential profiles. From the constraint on the scaled radius vs.\\ time curve and the physical radius of the remnant, we obtain a relation between the ejected mass and the density of the ambient medium. Assuming a Chandrasekhar mass for the ejecta, the density is $\\sim$4--$5\\times10^{-2}$ cm$^{-3}$ for the constant density ejecta profile of \\citet{tm}, and $\\sim$0.7--$1\\times10^{-2}$ cm$^{-3}$ for the exponential density ejecta profile of \\citet{dc}. These estimates imply a low density environment and yield swept-up masses for hydrogen of 0.055--$0.068~{\\rm M{_\\sun}}$ or 0.0095--$0.013~{\\rm M{_\\sun}}$, respectively. The evolutionary models also can provide estimates of the explosion energy and age of the remnant, but they rely on knowing the shock velocity at the current epoch, which is only constrained to be $>$3600 km s$^{-1}$. The explosion energy is $>$$0.15\\times 10^{51}\\,(V_s/3600\\,\\rm km\\, s^{-1})^2$ erg (constant density) and $>$$0.09\\times 10^{51}\\,(V_s/3600\\,\\rm km\\, s^{-1})^2$ erg (exponential). The age is $<$$860\\,(V_s/3600\\,\\rm km\\, s^{-1})^{-1}$ yr (constant density) and $<$$670\\,(V_s/3600\\,\\rm km\\, s^{-1})^{-1}$ yr (exponential). \\subsection{Non-Thermal Continuum} \\e05 is believed to be situated in a low density environment. There is the dynamical argument just given, which constrains the density to $n_{\\rm H} \\lesssim0.05$ cm$^{-3}$. \\citet{tuohy} obtained a value of $n_{\\rm H} \\lesssim0.02$ cm$^{-3}$ through optical studies. In addition, the remarkable symmetry of the outer shell of this SNR suggests the ambient density is much lower than that surrounding other remnants (given the same level of density fluctuations). For example, we do not see breaks in the rim or an irregular shape in \\e05 as we do in Tycho \\citep{hwang02}, which has an ambient density of $n_{\\rm H} \\sim0.5$ cm$^{-3}$ \\citep{hughes}. Our Case H deprojection yields a post-shock density of $3.6~{\\rm cm}^{-3}$ for hydrogen, which in turn yields a swept-up hydrogen mass of 7.4 ${\\rm M{_\\sun}}$. This density, converted to its preshock value of $\\sim$1 cm$^{-3}$ assuming the strong shock compression factor of 4, is much too high to be consistent with the previous arguments. If we take the ambient density to be $n_{\\rm H}=0.05$ cm$^{-3}$, then the level of hydrogen thermal continuum produced would be approximately $3\\times 10^{-3}$ times lower than the {\\em actual} continuum level observed in the spectrum. Therefore we are led to conclude that there is a significant non-thermal component of continuum emission in this SNR. In Table 4 we show the masses of H, O, Si, S, and Fe in the two outermost shells, derived from the densities obtained by our deprojection analysis under spherical geometry. Although the masses derived from the non-thermal case are too high to be consistent with SNe Ia (0.76 M$_\\sun$ for Si), this is a limiting case since it assumes {\\em all} the electrons come from the partially ionized metals. Therefore, it is an extreme upper estimate to the mass. If even a small amount of hydrogen were included with the non-thermal emission, the densities, and hence masses, of the metals would decrease. In addition, the clumpy nature of the SNR inferred from our image fits (\\S 3.2) suggests that the apparently diffuse emission (from the shell component in the image fits) may arise from a smaller volume than expected if it were uniformly spread throughout. The knots modeled in the image fits produce $\\sim$20\\% of the X-ray events from the shell in the northeast half of the remnant, while occupying only $\\sim$2\\% of the shell volume of the diffuse emission. If we assume that the diffuse component is made up of similar, though unresolved, knots with the same X-ray events-to-volume ratio as those fit directly, we find that only $\\sim$10\\% of the diffuse emission volume need be occupied by knots to produce the same emission. This 10\\% filling factor translates to a $\\sim$30\\% reduction in the derived masses, bringing the values for the non-thermal case to a more reasonable level: about 0.2 M$_\\sun$ for Si. Several other young, shell-like SNRs show evidence for non-thermal emission \\citep{koy2,allen,slane}, the most well-known being SN1006 \\citep{koy}. The shell of SN1006 is dominated by non-thermal emission, showing a featureless spectrum that can be described by a power-law with a photon index of 2.95 \\citep{koy}. The synchrotron cut-off model was also applied to SN1006 and a value of $6\\times10^{16}$ Hz for the cut-off frequency was obtained \\citep{reykeo}. For \\e05 we find a higher power-law photon index ($\\alpha_p\\sim3.3$) and a lower synchrotron cut-off frequency ($1.6\\times10^{16}$ Hz). Both of these results are consistent with \\e05 showing a steeper non-thermal spectrum in the 0.5--7 keV X-ray band than SN1006. As to the overall intensity of non-thermal emission, these two remnants are quite similar as well, with \\e05 being about a factor of three more luminous than SN1006. Other SNRs, such as Tycho, Kepler, and Cas A, are not dominated by non-thermal emission but are well-fit by a power-law at energies $\\gtrsim$10 keV (Allen et al. 1999). In the case of \\e05, we get as good, if not a better, fit for the global spectrum with a non-thermal continuum. Thus a non-thermal origin for the continuum emission from \\e05 is plausible. We also applied the minimum-energy condition for the synchrotron emission (Longair 1994, pp.~292-296) from this remnant, assuming equal energy densities for the protons and electrons and a volume filling factor of one. By plotting the extrapolated radio power-law against the cut-off power-law determined by the synchrotron cut-off model, we found that the latter begins to deviate from a straight power-law at $\\sim$10$^{14}$ Hz. We used this value as the maximum frequency in our calculation and note that varying this value did not change our results greatly. We obtained a minimum energy of $10^{48}$ ergs and magnetic field of 60 $\\mu$G. The energy requirements on this component are safely below the available energy of $10^{51}$ ergs, but our magnetic field value is large. However, the minimum energy condition (effectively equipartition between particles and the magnetic field) is unlikely to be applicable, as shown by Dyer et al.~(2001) for SN1006 from comparing TeV gamma-ray and non-thermal X-ray emission. If we assume a magnetic field of $\\sim$ 10 $\\mu$G as they find, then the total energy required to explain the non-thermal emission in 0509-67.5 increases to $10^{49}$ ergs, still well below the available energy. Using $\\sim$ 10 $\\mu$G for the magnetic field along with the fitted value of the roll-off frequency from the synchrotron cut-off spectral model, we find the maximum energy of electrons to be $\\sim$ 20 TeV. The preceding arguments strongly support the picture that the forward shock of \\e05 is accelerating particles to relativistic energies. \\e05 is thus the first Magellanic Cloud SNR for which such an energetic cosmic ray component has been securely detected. One task that remains to be done to make our work fully consistent is to explore the effects of modifications to the remnant dynamics due to efficient particle acceleration. As the fraction of the shock front's kinetic energy being diverted to the acceleration of relativistic particles grows, the thicknesses of the forward and reverse shocked regions tend to decrease \\citep{dec2}. Such changes could affect our estimate of the ambient medium density based on the thickness of the X-ray emitting shell. In the absence of a specific model for \\e05, it is not possible to estimate a correction factor for the ambient density. Although beyond the scope of this work, development of such a model would be a valuable next step toward determining the efficiency of cosmic ray acceleration in \\e05." }, "0403/astro-ph0403669_arXiv.txt": { "abstract": "We present {\\it Chandra} and {\\it XMM-Newton} X-ray data of NGC\\,5253, a local starbursting dwarf elliptical galaxy, in the early stages of a starburst episode. Contributions to the X-ray emission come from discrete point sources and extended diffuse emission, in the form of what appear to be multiple superbubbles, and smaller bubbles probably associated with individual star clusters. {\\it Chandra} detects 17 sources within the optical extent of NGC\\,5253 down to a completeness level corresponding to a luminosity of $1.5 \\times 10^{37}\\ergs$. The slope of the point source X-ray luminosity function is $-0.54 \\pm^{0.21}_{0.16}$, similar to that of other nearby dwarf starburst galaxies. Several different types of source are detected within the galaxy, including X-ray binaries and the emission associated with star-clusters. Comparison of the diffuse X-ray emission with the observed H$\\alpha$ emission shows similarities in their extent. The best spectral fit to the diffuse emission is obtained with an absorbed, two temperature model giving temperatures for the two gas components of $\\sim 0.24\\kev$ and $\\sim 0.75\\kev$. The derived parameters of the diffuse X-ray emitting gas are as follows: a total mass of $\\sim 1.4 \\times 10^{6} f^{1/2}\\msun$, where $f$ is the volume filling factor of the X-ray emitting gas, and a total thermal energy content for the hot X-ray emitting gas of $\\sim 3.4 \\times 10^{54}f^{1/2}$~erg. The pressure in the diffuse gas is $P/k\\sim 10^{6} f^{-1/2}$~K~cm$^{-3}$. We find that these values are broadly commensurate with the mass and energy injection from the starburst population. Analysis of the kinematics of the starburst region suggest that the stellar ejecta contained within it can escape the gravitational potential well of the galaxy, and pollute the surrounding IGM. ", "introduction": "NGC\\,5253 is a starbursting dwarf elliptical galaxy which lies at a distance of $\\sim 3.15$~Mpc (Freedman \\etal\\ 2001) and with an inclination of 67$^{\\circ}$ (the mean value quoted by the Lyon/Meudon Extra-galactic Database [LEDA]). The signature of Wolf-Rayet stars have been detected in the nucleus of the galaxy (Schaerer \\etal\\ 1997, Walsh \\& Roy 1987) implying that this is a young starburst and as such allows an opportunity for the study of the starburst phenomenon in its earlier stages of development and its effect on galaxy evolution in the local Universe. In fact, Rieke, Lebofsky \\& Walker (1988) went so far as to classify it as one of the youngest starbursts known. Dwarf galaxies, as the basic building blocks in the hierarchical merging cosmology scenario, are likely to have harboured the earliest sites of star-formation in the Universe and so their study in the local Universe can give insight into the evolution of such objects at high redshift. Observations of local edge-on starburst galaxies (Strickland \\etal\\ 2000; Weaver 2001) are presenting a picture of kpc-scale, soft X-ray emitting, bipolar outflows in the form of galactic winds transporting mass, newly synthesised heavy elements and energy into the intergalactic medium (IGM). A similar outflow is seen in the dwarf starburst galaxy NGC\\,1569 (Martin, Kobulnicky \\& Heckman 2002), while the dwarf NGC\\,4449 shows what may be the beginnings of a galactic wind emerging from an extended superbubble (Summers \\etal\\ 2003). The same situation seems to be the case in the NGC\\,3077 dwarf (Ott, Martin \\& Walter 2003). These winds result from the pressure driven outflows along these galaxy's minor axis produced from the efficient thermalization of the mechanical energy from the supernovae (SN) explosions and stellar winds of the massive stars in their OB associations and super star-clusters (SSC). NGC\\,5253 being both inclined to our line-of-sight and young presents a less clearly observable picture. Its youth means that it is unlikely to have developed a superwind and the diffuse X-ray emission observed is much more likely to be associated with multiple superbubbles around its OB associations and SSC, as suggested from earlier {\\it ROSAT} observations of this galaxy (Strickland \\& Stevens 1999). Observations of NGC\\,5253 at radio wavelengths have shown it to have a very flat centimetre-wavelength continuum which is indicative of thermal emission from H{\\small II} regions, with only low levels of synchrotron emission from supernova remnants (Beck \\etal\\ 1996; Turner, Ho \\& Beck 1998). Ferrarese \\etal\\ (2000) quote a low metallicity of $0.2 Z_{\\odot}$, based on the work of Webster \\& Smith (1983). This figure suggests that a large amount of metal enrichment has not occurred in the central H{\\small II} regions, and may be further evidence of the youth of the starburst region. The total and H{\\small I} masses for the galaxy are calculated to be $6.4 \\times 10^{8}\\msun$ and $8.3 \\times 10^{7}\\msun$ from 21~cm observations (Reif \\etal\\ 1982), corrected for the difference in assumed distance to NGC\\,5253. Later VLA 21~cm observations show NGC\\,5253 to be peculiar, in so much as its neutral hydrogen appears to rotate about the optical major axis of the galaxy (Kobulnicky \\& Skillman 1995). However, CO observations (Turner, Beck \\& Hurt 1997; Meier, Turner \\& Beck 2002) detect sources which are coincident with the dust lane seen to the SE of the nucleus of the galaxy in optical images, and these CO clouds appear to be infalling into NGC\\,5253. Consequently, the dynamical situation in NGC\\,5253, rotation or infall, is not completely clear. Infrared emission is detected from what seems to be a highly obscured, massive ($10^{5} -10^{6}$ stars), small ($\\sim$ a few pc diameter) and very young globular cluster in the central starburst region of the galaxy (Gorjian, Turner \\& Beck 2001). This source contributes $\\sim 50\\%$ of the total observed infrared luminosity of the galaxy. A summary of some of the properties of NGC\\,5253 is given in Table~1, along with estimates of the current star-formation rate of the galaxy estimated from far-infrared and H$\\alpha$ fluxes. Caldwell \\& Phillips (1989), using broadband multicolour (UBV) CCD images, narrow-band H$\\alpha$ and long-slit Ca{\\small II} triplet spectra, determined that NGC\\,5253 had undergone an increased rate of star-formation throughout the galaxy about $10^{8} - 10^{9}$~yr ago, as suggested by the presence of over 100 star clusters of a similar age lying outside the nuclear region. They also established that the intense star-formation now occurring in the nuclear region of the galaxy began around 10~Myr ago, giving an upper limit for the age of superbubbles associated with it. More recently, Tremonti et al. (2001) analysed several clusters in NGC\\,5253, determining ages in the range of 1 to 8~Myr. Calzetti et al. (1997) also noted two other clusters with older ages, determined to be between 10 and 50~Myr. The two very bright SN explosions seen in this galaxy in 1895 and 1972 are amongst the apparently brightest recorded extra-galactic SNe (Ardeberg \\& de Groot 1973) and may be associated with the large scale star-formation, although neither was particularly centrally located. The Fabry-Perot H$\\alpha$ images of Marlowe \\etal\\ (1995) show that the ionized gas in the galaxy has a complex distribution, consisting of a bright central region extending $\\sim 30{''}$ from the centre in a N-S direction, embedded within a system of loops and filaments, with an overall extent of $\\sim 2{'}$. To the NW and WSW of the centre are 2 large superbubbles with diameters $\\sim 1{'}$, expanding at $\\sim 35\\kms$, while to the east side of the galaxy, a pair of radially oriented filaments are detected. This latter observation is suggestive of the rupture of a superbubble blown by the stellar activity in the central regions. NGC\\,5253 has previously been observed at X-ray wavelengths with the {\\it Einstein} IPC (Fabbiano, Kim \\& Trinchieri 1992) and both the {\\it ROSAT} PSPC (Martin \\& Kennicutt 1995; Stevens \\& Strickland 1998) and HRI (Strickland \\& Stevens 1999). The {\\it Einstein} IPC and {\\it ROSAT} PSPC results showed soft thermal X-ray emission from what appeared to be an extended source, suggestive of a luminous superbubble. The better spatial resolution of the {\\it ROSAT} HRI instrument detected a complex of at least five X-ray sources, that could be young superbubbles blown by individual young star clusters in the starburst region. The total X-ray luminosity determined from these observations was similar and lies in the range of $(2.4 - 4.1) \\times 10^{38}\\ergs$, depending on the value adopted for the absorbing column density. In Section 2 we describe the {\\it Chandra} and {\\it XMM-Newton} observations. The X-ray emission from the whole galaxy, the point sources and the diffuse emission are discussed in Section~3. Section~4 contains a more general discussion of the morphology of the X-ray emission and its relationship to emission from other wavebands, along with a determination of the dynamics of the expanding superbubbles of the central region and their effect on NGC\\,5253. Our main conclusions are summarised in Section~5. \\begin{figure*} \\vspace{13cm} \\special{psfile=n5253_fig1_alt.ps hoffset=-40 voffset=398 hscale=70 vscale=70 angle=-90} \\caption{Low resolution (smoothed using a Gaussian with FWHM of 4 pixels $\\sim 2^{''}$), background subtracted {\\it Chandra} image, in the $0.3 - 8.0\\kev$ energy band, showing the full ACIS-S3 chip field of view, marked with the 31 point sources (Table~2). The image is an inverted log greyscale with the flux density ranging from $7.63 \\times 10^{-14}\\ergs$~cm$^{-2}$~arcmin$^{-2}$ to $9.72 \\times 10^{-11}\\ergs$~cm$^{-2}$~arcmin$^{-2}$. The $D_{25}$ ellipse is also shown. North is to the top and East is to the left.} \\label{sources} \\end{figure*} \\begin{table} \\caption{Details of NGC\\,5253. The distance is from Freedman \\etal\\ (2001), other data is from the LEDA and NED databases. The infrared luminosity $L_{IR}$ is calculated using the 12, 25, 60 and $100\\mu$m IRAS fluxes (Sanders \\& Mirabel 1996) and the H$\\alpha$ luminosity is from Marlowe \\etal\\ (1997, corrected for distance). The corresponding star-formation rates (SFR) for both the IR and H$\\alpha$ luminosities are calculated using the conversion formulae in Kennicutt (1998).} \\begin{center} \\begin{tabular}{ll} \\hline Parameter & Value \\\\ \\hline Classification\\ \\ \\ \\ \\ & Im pec, H{\\small II} \\\\ Distance & 3.15Mpc \\\\ RA (J2000) & 12 39 56.0 \\\\ Dec (J2000) &$-31$ 38 36\\\\ $D_{25}$ ellipse & $4.8' \\times 1.9'$\\\\ Position angle & $45^\\circ$\\\\ $m_B$ & 10.75\\\\ $B-V$ & 0.43\\\\ $L_{IR}$ & $4.5\\times 10^{42}\\ergs$\\\\ $L_{\\rm {H}\\alpha}$ & $2.6\\times 10^{40}\\ergs$\\\\ SFR rate (IR) & $0.2\\msunyr$\\\\ SFR rate (H$\\alpha$) & $0.2\\msunyr$\\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\begin{table} \\begin{center} \\caption{Positions and count rates of the 31 sources detected in the NGC\\,5253 {\\it Chandra} S3 chip data. Column 1 is the source number ordered in increasing R.A.. Columns 2 and 3 give the R.A. and Dec. of each source and Column 4 lists their background subtracted count rates.} \\begin{tabular}{|c|c|c|c|} \\hline Source & RA (h m s) & Dec ($^{\\circ} \\; {'}\\; {''}$) & Count Rate \\\\ & & & ($\\times 10^{-3}$ cts s$^{-1}$) \\\\ \\hline 1 & 13 39 37.73 & -31 40 15.8 & $0.77\\pm0.15$ \\\\ 2 & 13 39 39.16 & -31 36 06.9 & $0.32\\pm 0.09$ \\\\ 3 & 13 39 39.17 & -31 36 09.2 & $0.42\\pm 0.10$ \\\\ 4 & 13 39 43.15 & -31 39 20.2 & $1.07\\pm 0.16$\\\\ 5 & 13 39 45.12 & -31 36 58.2 & $0.37\\pm 0.10$\\\\ 6 & 13 39 45.53 & -31 42 06.1 & $0.39\\pm 0.10$\\\\ 7 & 13 39 46.69 & -31 42 28.0 & $0.43\\pm 0.11$\\\\ 8 & 13 39 47.27 & -31 41 40.7 & $1.95\\pm 0.22$\\\\ 9 & 13 39 50.66 & -31 39 20.7 & $1.50\\pm 0.18$\\\\ 10 & 13 39 51.17 & -31 38 40.8 & $0.32\\pm 0.09$\\\\ 11 & 13 39 51.62 & -31 40 16.2 & $0.48\\pm 0.11$\\\\ 12 & 13 39 51.96 & -31 41 43.9 & $4.03\\pm 0.30$\\\\ 13 & 13 39 52.07 & -31 38 59.7 & $0.19\\pm 0.07$\\\\ 14 & 13 39 53.36 & -31 43 43.9 & $1.42\\pm 0.20$\\\\ 15 & 13 39 53.80 & -31 38 32.7 & $0.66\\pm 0.13$\\\\ 16 & 13 39 55.39 & -31 38 32.8 & $0.80\\pm 0.17$\\\\ 17 & 13 39 55.69 & -31 38 31.5 & $1.02\\pm 0.20$\\\\ 18 & 13 39 55.85 & -31 38 22.8 & $1.03\\pm 0.20$\\\\ 19 & 13 39 56.04 & -31 38 24.8 & $1.43\\pm 0.23$\\\\ 20 & 13 39 56.15 & -31 38 35.9 & $0.58\\pm 0.16$\\\\ 21 & 13 39 56.35 & -31 38 20.5 & $8.22\\pm 0.43$\\\\ 22 & 13 39 56.45 & -31 38 25.8 & $1.53\\pm 0.21$\\\\ 23 & 13 39 56.67 & -31 42 45.3 & $1.57\\pm 0.21$\\\\ 24 & 13 39 57.43 & -31 37 32.5 & $1.50\\pm 0.06$\\\\ 25 & 13 39 59.27 & -31 37 50.2 & $0.33\\pm 0.09$\\\\ 26 & 13 40 02.19 & -31 37 19.1 & $0.96\\pm 0.15$\\\\ 27 & 13 40 03.40 & -31 37 14.5 & $0.78\\pm 0.13$\\\\ 28 & 13 40 03.84 & -31 44 03.7 & $1.25\\pm 0.21$\\\\ 29 & 13 40 04.29 & -31 37 30.3 & $0.41\\pm 0.10$\\\\ 30 & 13 40 04.97 & -31 42 03.0 & $0.24\\pm 0.08$\\\\ 31 & 13 40 10.81 & -31 40 33.2 & $0.29\\pm0.09$\\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\begin{figure*} \\vspace{13cm} \\special{psfile=n5253_fig2.ps hoffset=-20 voffset=375 hscale=70 vscale=70 angle=-90} \\caption{Low resolution (smoothed using a Gaussian with FWHM of 2 pixels $\\sim 8^{''}$), {\\it XMM-Newton}, background subtracted and exposure corrected, mosaiced image of the data from the EPIC MOS1, MOS2 and PN cameras, in the $0.3 - 8.0\\kev$ energy band. The image is an inverted log greyscale with the flux density ranging from $5.2 \\times 10^{-14}\\ergs$~cm$^{-2}$~arcmin$^{-2}$ to $2.1 \\times 10^{-12}\\ergs$~cm$^{-2}$~arcmin$^{-2}$, with North to the top and East to the left. The $D_{25}$ ellipse is also shown. The X-ray emission is extended towards the WSW.} \\label{mosaic} \\end{figure*} ", "conclusions": "In summary, we have presented an X-ray analysis of the important dwarf elliptical, starburst galaxy NGC\\,5253 using data from both the {\\it Chandra} and {\\it XMM-Newton} satellites. We detected X-ray emission from 31 discrete point sources in the {\\it Chandra} ACIS-S3 chip data with 17 of these lying within the optical extent of the galaxy, as measured by the $D_{25}$ ellipse. Some of these can be clearly identified as X-ray binaries while others seem to be associated with the emission from super star-clusters rather than individual point sources, as they are extended compared to the {\\it Chandra} PSF. The 5 sources in this category (sources $16 - 20$) all lie in the central region of the galaxy and are associated (particularly in the case of sources 17, 18 and 19) with the most intense regions of emission seen on the H$\\alpha$ image. The unabsorbed luminosities of these sources in the $0.3 - 8.0\\kev$ energy band, range from $(0.25 - 1.50) \\times 10^{38}\\ergs$. The majority of the X-ray emission is confined to the central region of the galaxy where the intense, current burst of star-formation is occurring. The extent of the emission also correlates well with that of the H$\\alpha$ emission and both are confined to a region of diameter $\\sim 1$~kpc in the centre of the galaxy. Both of these emissions are embedded within a spherical distribution of H{\\small I} that extends out to $\\sim 2$~kpc from the centre. The diffuse X-ray emission in NGC\\,5253 requires two thermal and one power-law component to best fit its spectrum, suggesting that the hot gas is a multi-phase environment with unresolved point sources, probably lower luminosity X-ray binaries, embedded in it. The fitted gas temperatures are $0.24 \\pm 0.01\\kev$ and $0.75 \\pm 0.05\\kev$ for the soft and medium components respectively and their respective absorption corrected luminosities in the $0.37 - 6.0\\kev$ band are $(2.19 \\pm^{0.16}_{0.15}) \\times 10^{38}\\ergs$ and $(1.75 \\pm^{0.15}_{0.14}) \\times 10^{38}\\ergs$. This emission is seen to be extended in several directions in the {\\it Chandra} data but this is not confirmed by the {\\it XMM-Newton} data. However, the extent is not as great as the $\\sim 1$~kpc shells reported to the NW and WSW by Marlowe \\etal\\ (1995), in the H$\\alpha$ emission. These shells are expanding with velocities of $\\sim 35\\kms$ and hence have required $\\sim$10~Myr to have evolved. NGC\\,5253 is proving to be a rather enigmatic object though, as our analysis of the dynamics of the hot gas are not consistent with the expansion velocity and age of these H$\\alpha$ shells. The hot X-ray emitting gas has a total thermal energy content of $3.4 \\times 10^{54}$~erg and a total mass of $1.4 \\times 10^{6}\\msun$. From values in the literature, we assume an age for the starburst responsible for this hot gas of 5~Myr and a value for the density of the ambient ISM of $n_{0} = 5$~cm$^{-3}$. The radius of the region in which most of the diffuse X-ray and H$\\alpha$ emission is confined is $0.53$~kpc. Application of standard superbubble models to NGC\\,5253's starburst, assuming the age and density above, result in a predicted radius and expansion velocity for a single superbubble of $\\sim 400$~pc and $\\sim 50\\kms$, figures which are comparable to those observed. The main problem here would seem to lie in the fact that no clear bipolar outflow is observed in NGC\\,5253 and no allowance has been made for the presence of multiple superbubbles in the central region. If the central starburst is young, the bubbles blown around individual star clusters will overlap and impede each others expansion confining the emission to a smaller region than expected and reducing the observed expansion rate. The H$\\alpha$ shells, reported by Marlowe \\etal\\ (1995), to the West could be the result of the superbubbles closest to the edge of the central region being able to expand more freely. The lower velocity seen could result from less stars contributing to the energy input in these regions and/or the outflow having experienced some recent deceleration on encountering higher density material such as the dark clouds reported in the West of the galaxy (Hunter 1982). However, our radial surface brightness profile does not show any detectable X-ray emission beyond $\\sim 0.5$~kpc. NGC\\,5253 is a very complex object, which shows that standard superbubble models are difficult to apply to dwarf starbursts galaxies. However, the energy injection rate into the galaxy would seem sufficient to allow the expanding hot gas to escape the gravitational potential well of NGC\\,5253 and its relatively small ($\\sim 1.8$~kpc along the minor axis) H{\\small I} halo would also offer little resistance to this. Thus, it seems likely then that NGC\\,5253 will lose metal-enriched material, mass and energy as a result of its current bout of star-formation." }, "0403/astro-ph0403496_arXiv.txt": { "abstract": "We report the strength of seed magnetic flux of accretion disk surrounding the PopIII stars. The magnetic field in accretion disk might play an important role in the transport of angular momentum because of the turbulence induced by Magneto-Rotational Instability (MRI). On the other hand, since the primordial star-forming clouds contain no heavy elements and grains, they experience much different thermal history and different dissipation history of the magnetic field in the course of their gravitational contraction, from those in the present-day star-forming molecular clouds. In order to assess the magnetic field strength in the accretion disk of PopIII stars, we calculate the thermal history of the primordial collapsing clouds, and investigate the coupling of magnetic field with primordial gas. As a result, we find that the magnetic field strongly couple with primordial gas cloud throughout the collapse, i.e. the magnetic field are frozen to the gas as far as initial field strength satisfies $ B\\lesssim 10^{-5}(n_{\\rm H}/10^3~{\\rm cm^{-3}})^{0.55}~{\\rm G}$. ", "introduction": "PopIII stars are considered to have very important impacts on the thermal history and the chemical evolution of the universe. Recent observations of the polarization of Cosmic Microwave Background (CMB) photons by WMAP(Wilkinson Microwave Anisotropy Probe) \\citep{Kogut03} revealed that the optical depth of the universe by Thomson scattering is fairly large ($\\tau = 0.17 \\pm 0.04$). This result tells that reionization epoch is earlier ($z = 17 \\pm 5$) than expected from the observations on Gunn-Peterson trough \\citep{Becker01,Fan02,White03}. On the other hand, the results of theoretical calculations of structure formation \\citep{Sokasian03} indicate that such early reionization seems to be difficult solely by the Pop II stars, but it might be possible if the ionizing photons from PopIII stars with top heavy initial mass function (IMF) is taken into account. Thus, the mass (or IMF) of PopIII star have great significance on the reionization epoch of the universe. In addition, some amount of metals are found at high redshift intergalactic matter by the observation of high redshift QSO absorption line systems\\citep{songaila01,ma02,vladilo02}. This means significant fraction of baryons are already processed in the stars by $z=5$ \\citep{songaila01}. PopIII stars also should be responsible for such metal pollution of the early universe, and the abundance pattern of heavy elements depends on the mass of PopIII stars. Thus, typical mass (or IMF) of the PopIII stars is a key quantity also for the chemical evolution of the universe. In order to assess the typical mass of PopIII stars, typical scale of the prestellar core (or fragments of primordial gas) should be evaluated as a first step. Several authors have studied on this issue by simple one-zone approach before middle of '90s \\citep{MST69,Hut,Carl,PSS83,SUN96,USNY96,Puy}. Recently, multi-dimensional numerical simulations of fragmentation of primordial gas have been performed intensively by several authors\\citep{nakamura99,bromm99,abel00,nakamura01,bromm02}, and they find that the mass of PopIII prestellar core is quite massive ($\\sim 10^{3-4}M_\\odot$). Further evolution of prestellar core was investigated by \\cite{omukai98}, and it is found that the collapse proceeds in a run-away fashion and converges to Larson-Penston similarity solution \\citep{Larson69,Penston69} with $\\gamma\\simeq 1.1$. They also find the mass accretion rate is also very large compared to the present-day forming stars, although spherical symmetry is assumed in their numerical calculation. In reality, however, prestellar cores have some amount of angular momentum, which prevent the mass from accreting onto the protostars. Consequently, accretion disks surrounding the protostars are expected, and some mechanisms that transport their angular momentum are required in order to enable the mass accretion onto protostars. There are a few possibilities of the angular momentum transport such as gravitational torque by the nonaxisymmetric structures in the accretion disk, the interaction among the fragments of the disk \\citep{stone00,bodenheimer00}, and the turbulent viscosity triggered by Magneto-Rotational Instability (MRI) \\citep{hawley92,sano98,sano01}. All of these mechanisms are regarded to be important for present-day star formation. Thus, it is worth to investigate these processes for primordial case. Among these possibilities, we concentrate on the last one, the MRI induced turbulence. In order to activate MRI, the initial magnetic field strength in the accretion disk should be larger than a critical value. Otherwise the magnetic field is dissipated before the field is amplified by MRI \\citep{TM04,TB04}. Thus, it is important to assess the ``initial'' magnetic field strength in the accretion disk. As was discussed by \\citet{nakano86} for preset-day case, magnetic field could be dissipated while prestellar core collapses. However, the dissipation processes strongly depend on the components of the gas as well as the the temperature evolution in the course of the collapse. On the other hand, the temperature of the collapsing primordial gas is much higher than that of the present-day case \\citep{omukai00}, which might bring about different dissipation history of the magnetic field. In this paper, we investigate the dissipation of magnetic field in collapsing primordial gas cloud. Consequently, we obtain the initial magnetic field strength in the accretion disk of PopIII stars. In the next section, formulations are described. In \\S\\ref{results}, results of our calculations are shown, and the importance of MRI in PopIII star formation is discussed in \\S\\ref{discussion}. Final section is devoted to summary. ", "conclusions": "\\label{discussion} The condition for which the MRI can grow at the PopIII accretion disk is investigated by \\citet{TB04}. This condition requires that the MRI growth timescale is shorter than the diffusion timescale. Thus, there is the minimum field strength in the disk for the MRI to be driven. According to \\citet{TB04}, this minimum field strength in the disk becomes \\begin{eqnarray} B&=&1.1\\times 10^{-4}{\\rm G} \\left(\\frac{m_*}{10\\ M_\\odot}\\right)^{1/4} \\left(\\frac{T}{10^4\\ {\\rm K}}\\right)^{-3/4} \\left(\\frac{\\ln\\Lambda}{10}\\right)^{1/2}\\nonumber\\\\ &&\\times\\left(\\frac{\\rho_{\\rm disk}} {5\\times 10^{-10}\\ {\\rm g\\;cm^{-3}}}\\right)^{1/2} \\left(\\frac{r}{600\\ R_\\odot}\\right)^{-3/4}, \\label{eq:discon} \\end{eqnarray} where $m_*$ is the stellar mass, $r$ is the radius from the cloud center, $\\rho_{\\rm disk}$ denotes the density of the accretion disk and $\\ln \\Lambda$ is the Coulomb logarithm. Characteristic values of these parameters are taken from Figure 3 in \\citet{TB04}. Hence, it is concluded that if the initial field strength in prestellar core is at least $\\gtrsim 10^{-10}\\ {\\rm G}$ at $n_{\\rm H}=10^3\\ {\\rm cm^{-3}}$, the transport of angular momentum could be driven by the turbulence due to the MRI in the accretion disk. In fact, there are the models for the generation of initial seed magnetic field in the universe \\citep{pudritz89,kulsrud97,widrow02,langer03}. Among these models, \\citet{langer03} propose the generation mechanism of magnetic field based on the radiation force around very luminous objects such as QSOs. According to their study, the magnetic field $\\sim 10^{-11}-10^{-12}\\ {\\rm G}$ is generated in the inter galactic matter. Consequently, field strength is amplified to $\\sim 10^{-7}-10^{-8}\\ {\\rm G}$ when the clouds collapses to $n_{\\rm H}=10^3\\ {\\rm cm^{-3}}$, which is the initial condition of our analysis. Thus, in this case, the MRI can be driven and it could be the possible mechanism of angular momentum transport. On the other hand, most of the other seed field generation mechanism predict $B \\lesssim 10^{-19}\\ {\\rm G}$, which is too small to drive MRI. Thus, MRI may not be important for the formation of very first stars, since the number of sources which provide anisotropic radiation field should be very small at the epoch of very first stars." }, "0403/astro-ph0403482_arXiv.txt": { "abstract": "We report the discovery of only the fourth massive WO star to be found in the Milky Way, and only the seventh identified within the Local Group. This has resulted from the first observations made in a programme of follow-up spectroscopy of candidate emission line stars from the AAO/UK Schmidt Southern Galactic Plane H$\\alpha$ Survey. The optical spectrum of this star, to become WR~93b in the Catalogue of Galactic Wolf-Rayet stars, is presented and described. WR~93b is classified as WO3 and is shown to be highly reddened ($E_{B-V} = 2.1\\pm0.1$). A recombination line analysis of the emission lines yields the abundance ratios C/He = 0.95 and O/He = 0.13 (by number). Comparisons at near infrared wavelengths of reddening corrected photometry between WR~93b and both of Sand~2 (WO3, $D = 49$~kpc) and Sand~5 (WO2, $D = 1.75$~kpc) yields a consistent distance to WR~93b of 3.4~kpc. Positioned at Galactic co-ordinates $\\ell = 353.27^{\\rm o}$, $b = -0.85^{\\rm o}$, the star is most likely located in the Scutum-Crux Arm of the inner Milky Way. We note that none of the four Galactic WO stars lies significantly beyond the Solar Circle (with two well inside). Estimation of the wind terminal velocity in WR~93b at 5750~km~s$^{-1}$ makes this star the current wind speed record holder among all non-degenerate stars. ", "introduction": "In this paper we present the discovery of a particularly rare type of emission line star made in the initial phase of a long-term spectroscopy programme which aims to confirm and provide preliminary classification of spatially-unresolved line-excess objects contained within the AAO/UK Schmidt Southern Galactic Plane H$\\alpha$ Survey (Parker et al 2003). The star in question has been revealed as only the fourth massive WO star to be found in the Galaxy -- bringing the total identified within the Local Group to just seven. When it is assimilated into the next revision of the catalogue of massive Wolf-Rayet stars in the Galaxy, compiled and maintained by van der Hucht (2001), its position in the sky will earn it the designation WR 93b. Hereafter we refer to this newly discovered star by this name. The WO stars are the most chemically extreme Wolf-Rayet (WR) stars, whose spectra are dominated by high-excitation oxygen and carbon lines. Objects in this elite group of stars are viewed as plausible progenitors for extreme, chemically-peculiar Type Ib/c supernovae (Woosley, Heger \\& Weaver 2002) and some GRBs (e.g. Schaefer et al 2003). The history of this presently rare type of object begins with the last generation of galactic emission line surveys: Sanduleak (1971) presented a list of five WR stars, two in the Magellanic Clouds, which had strong O~{\\sc vi} 3811,34~\\AA\\ emission features that had previously been found only amongst planetary nebula nuclei. Barlow \\& Hummer (1982) proposed that one of these five stars was indeed a PN central star but that the other four corresponded to an advanced stage of evolution of massive stars, beyond the WC phase, and classified them as WO Wolf-Rayet stars. Since this time, two more have been added to this grouping (MS 4, by Smith, Shara \\& Moffat 1990 and DR~1 in the dwarf irregular galaxy IC 1613 by Kingsburgh \\& Barlow 1995). In general terms, discoveries of Galactic WR stars have been assisted greatly by optical objective-prism surveys (Stephenson \\& Sanduleak 1971, MacConnell \\& Sanduleak 1970) and narrow-band imaging surveys (most recently, Shara et al. 1999). The result we present here uses the latter technique in the red part of the spectrum, and serves as an encouragement that much remains to be discovered. In the next section we introduce the AAO/UK Schmidt Southern Galactic Plane H$\\alpha$ Survey and describe the data it contains relevant to the newly-discovered WO star and its locale. We then describe how the WO star came to be included in our initial programme of follow-up spectroscopy using the AAO/UK Schmidt multi-fibre spectroscopy facility, 6dF. This is followed, in Section 3, by presentation of both the 6dF data obtained in May/June 2003 (\\S 3.1), and flux-calibrated long-slit spectra obtained at the William Herschel Telescope (WHT) in August 2003 (\\S 3.2). We are then in a position to classify the WO star and determine its reddening (\\S 4). We also present estimates of C/He and O/He abundance ratios deduced from the emission lines (\\S 5). Finally, by comparison with other WO stars at known distances, we derive the distance to WR~93b (\\S 6): at 3.4~kpc, and along a line of sight passing within a few degrees of the Galactic Centre, it seems likely the new WO star is associated with the Scutum-Crux spiral arm (Russeil 2003) -- without a doubt it is well inside the Solar Circle. We close in \\S 7 with comment on the 4-strong group of galactic WO stars and on the prospects for future discoveries. ", "conclusions": "\\begin{tabular}{lll} \\hline Parameter & value & comment\\\\ \\hline Position (RA,Dec) & 17 32 03.30 $-$35 04 32.5 & J2000, from SHS database \\\\ $R$, $I$ magnitudes & 14.7, 12.5 & from SHS database\\\\ $J$, $H$, $K$ magnitudes & 11.33$\\pm$0.03, 10.56$\\pm$0.04, 10.17$\\pm$0.04 & from 2MASS database \\\\ E($B-V$) & 2.1$\\pm$0.1 & $R=3.1$ gives A$_V \\simeq 6.5$ \\\\ spectral type & WO3 & criteria place WR 93b closer \\\\ & & to WO2/3 boundary than Sand~2 \\\\ wind terminal velocity & 5750 km s$^{-1}$ & \\\\ abundance ratios: & C/He $=$ 0.95 & by number \\\\ & O/He $=$ 0.13 & \\ `` \\ \\ `` \\\\ & (C+O)/He $=$ 1.08 & \\ `` \\ \\ `` \\\\ distance & 3.4$\\pm$0.3 kpc & \\\\ \\hline \\end{tabular} \\end{table*} Adopting a distance of 49~kpc to the LMC (see e.g.Gibson 2000), the difference of 5.7 magnitudes with respect to Sand~2 yields a distance of 3.5~kpc to WR~93b. Two recent estimates of the distance to Be~87, the cluster hosting Sand~5, are respectively 1.6~kpc (Massey, DeGioia-Eastwood \\& Waterhouse 2001) and 1.9~kpc (Kn\\\"odlseder et al 2002): we adopt the intermediate value of 1.75~kpc. The NIR magnitude difference of $-$1.43, combined with this distance then yields a distance to WR~93b of 3.4~kpc. This is very satisfying agreement. But, given the systematic uncertainty contained within the assumption of shared stellar parameters, we should anticipate an error on 3.4~kpc of up to 10 \\% (equivalent to $\\sim$0.2 offset in $M_V$). The Galactic co-ordinates of WR~93b are $\\ell = 353.27\\deg$, $b = -0.85\\deg$, placing it less than 7\\deg from the Galactic Centre line of sight, and outside the strips surveyed for Wolf-Rayet stars by Shara et al (1999). At a distance of 3.4~kpc the new WO star would be located very close to the structure Russeil (2003) identifies as the Scutum-Crux arm, well inside the long-established Sagittarius-Carina arm at $\\sim$2~kpc in the same direction. For a necessarily young object like a Wolf-Rayet star this is a perfectly acceptable place to be. Projected onto the plane of the sky, the nearest H~{\\sc ii} region (a signature of continuing star formation) listed by Russeil (2003, Table 3) is about half a degree away from WR~93b at $\\ell = 353.43\\deg$, $b = -0.368\\deg$. At 3.4~kpc this angular separation converts to a length of about 30~pc. The kinematic distance to this H{\\sc ii} region, which is `compact' and without a known optical counterpart, has been estimated from Caswell \\& Haynes' (1987) measurement of its radio recombination line radial velocity. In the framework adopted by Russeil (2003), wherein the Galactic Centre is 8.5~kpc away, the distance to this H~{\\sc ii} region is required to be $3.5^{+0.6}_{-0.7}$ kpc. This is consistent with our estimate of the distance to the WO star. If the Galactic Centre is taken to be 7.94$\\pm$0.42~kpc as determined recently by Eisenhauer et al (2003), this revises the kinematic distance downwards a little to 3.3~kpc -- leaving the quality of agreement unchanged." }, "0403/astro-ph0403161_arXiv.txt": { "abstract": "We present a redshift $z=6.535$ galaxy discovered by its \\lya\\ emission in a 9180\\AA\\ narrowband image from the Large Area Lyman Alpha (LALA) survey. The \\lya\\ line luminosity ($1.1\\times 10^{43} \\ergsec$) is among the largest known for star forming galaxies at $z\\approx 6.5$. The line shows the distinct asymmetry that is characteristic of high-redshift \\lya. The $2\\sigma$ lower bound on the observer-frame equivalent width is $>530$\\AA. This is hard to reconcile with a neutral intergalactic medium unless the \\lya\\ line is intrinsically strong {\\it and\\/} is emitted from its host galaxy with an intrinsic Doppler shift of several hundred $\\kms$. If the IGM is ionized, it corresponds to a rest frame equivalent width $> 40$\\AA\\ after correcting for \\lya\\ forest absorption. We also present complete spectroscopic followup of the remaining candidates with line flux $> 2\\times 10^{-17} \\ergcm2s$ in our $1200\\sqamin$ narrowband image. These include another galaxy with a strong emission line at $9136$\\AA\\ and no detected continuum flux, which however is most likely an \\oiii\\ source at $z=0.824$ based on a weak detection of the \\oiiib\\ line. ", "introduction": "Observational study of the redshift range $6 \\la z \\la 7$ is crucial for understanding the reionization of intergalactic hydrogen, which was the most recent major phase transition for most of the baryonic matter in the universe. While polarization of the microwave background measured by the Wilkinson Microwave Anisotropy Probe (WMAP) satellite indicates that substantial ionization had begun as early as $z \\sim 15$ (Spergel et al. 2003), both the opaque Gunn-Peterson troughs observed in the spectra of $z \\ga 6.3$ quasars (Becker et al. 2001; Fan et al. 2002) and the high temperature of the intergalactic medium at $z \\sim 4$ (Theuns et al. 2002; Hui \\& Haiman 2003) imply that a large part of the reionization was relatively recent, with substantial neutral gas lasting up to $z \\sim 6$. \\lya\\ emission has proven the most effective tool so far for identifying star forming galaxies at redshifts $z>6$; indeed, all but two of the $\\sim 6$ galaxies that have been spectroscopically confirmed at $z>6$ were found through their \\lya\\ line emission in either narrowband or spectroscopic searches (Hu et al 2002 [H02]; Kodaira et al 2003 [K03]; Cuby et al 2003; this work; Kneib et al 2004), and at least one of the continuum-selected sources is also a \\lya\\ emitter (Cuby et al 2003). We present the extension of the Large Area Lyman Alpha (LALA) survey to the $z\\approx 6.5$ window. We have obtained spectra for each of the three $z\\approx 6.5$ candidates that pass all photometric selection criteria. Two of these show strong emission lines in the narrowband filter bandpass. In one case, the line is identified as \\lya\\ at $z\\approx 6.535$ based on its asymmetric profile and the absence of other detected lines down to faint flux levels. The second object is identified as an \\oiii\\ source at $z\\approx 0.824$ based on a probable ($4\\sigma$) detection of the \\oiiib\\ line. ", "conclusions": "We have performed a narrowband search for \\lya\\ emitting galaxies at $z\\approx 6.5$ in the Bo\\\"{o}tes field of the Large Area Lyman Alpha survey, and have obtained spectra for all of our viable candidates. One source, \\gemwin, is confirmed as \\lya\\ at $z = 6.535$, based on an isolated, asymmetric emission line. The \\lya\\ luminosity of this source is large, $\\approx 1.1\\times 10^{43} \\ergsec$. Objects this bright in the line are rare, occurring at a rate $\\sim 1$ per $2\\times 10^5 \\hbox{comoving} \\Mpc^3$ (this work and K03). It is therefore likely to correspond to a high peak in the density distribution at redshift $z=6.535$. The equivalent width is also larger than most other $z\\approx 6.5$ galaxies, with a $2\\sigma$ lower bound of $>40$\\AA\\ (rest frame, corrected for IGM absorption). This is difficult to reconcile with a neutral intergalactic medium unless the \\lya\\ line is intrinsically strong {\\it and\\/} is emitted from its host galaxy with an intrinsic Doppler shift of several hundred $\\kms$." }, "0403/astro-ph0403027_arXiv.txt": { "abstract": "{ It is shown that if the self--gravitating shearing sheet, a model of a patch of a galactic disk, is embedded in a live dark halo, this has a strong effect on the dynamics of density waves in the sheet. I describe how the density waves and the halo interact via halo particles either on orbits in resonance with the wave or on non-resonant orbits. Contrary to expectation the presence of the halo leads to a very considerable enhancement of the amplitudes of the density waves in the shearing sheet. This effect appears to be the equivalent of the recently reported enhanced growth of bars in numerically simulated stellar disks embedded in live dark halos. Finally I discuss the transfer of linear momentum from a density wave in the sheet to the halo and show that it is mediated only by halo particles on resonant orbits. ", "introduction": "The shearing sheet (Goldreich \\& Lynden--Bell 1965, Julian \\& Toomre 1966) model has been developed as a tool to study the dynamics of galactic disks and is particularly well suited to describe theoretically the dynamical mechanisms responsible for the formation of spiral arms. For the sake of simplicity the model describes only the dynamics of a patch of a galactic disk. It is assumed to be infinitesimally thin and its radial size is assumed to be much smaller than the disk. Polar coordinates can be therefore rectified to pseudo-Cartesian coordinates and the velocity field of the differential rotation of the disk can be approximated by a linear shear flow. These simplifications allow an analytical treatment of the problem, which helps to clarify the underlying physical processes operating in the disk. In the present paper of this series I discuss the dynamical effects if the shearing sheet is immersed in a live dark halo. Dark halos are usually thought to stabilize galactic disks against non-axisymmetric instabilities. This was first proposed by Ostriker \\& Peebles (1973) on the basis of -- low--resolution -- numerical simulations. Their physical argument was that the presence of a dark halo reduces the destabilizing self--gravity of the disks. Doubts about an entirely passive role of dark halos were raised by Toomre (1977), but he (Toomre 1981) also pointed out that a dense core of a dark halo may cut the feed--back loop of the corotation amplifier of bars or spiral density waves and suppress thus their growth. Recent high-resolution numerical simulations by Debattista \\& Sellwood (2000) and Athanassoula (2002, 2003) have shown that quite the reverse, a {\\em destabilization} of disks immersed in dark halos, might be actually true. Athanassoula (2002) demonstrated clearly that much stronger bars grow in the simulations if the disk is embedded in a live dark halo instead of a static halo potential. This is attributed to angular momentum transfer from the bar to the halo via halo particles on resonant orbits. Angular momentum exchange between disk and halo has been addressed since the pioneering work of Weinberg (1985) in many studies theoretically or by numerical simulations and I refer to Athanassoula (2003) for an overview of the literature. Toomre (1981) has shown how the bar instability can be understood as an interference of spiral density waves in a resonance cavity between the corotation amplifier and an inner reflector (cf.~also Fuchs 2004). Thus it is to be expected that a live dark halo will be also responsive to spiral density waves and develop wakes, which I investigate here using the shearing sheet model. ", "conclusions": "If a self--gravitating shearing sheet is embedded in a live dark halo, the halo particles respond unexpectedly strong to density waves in the sheet. The interaction between the density waves and the halo particles is mediated both by halo particles on orbits in resonance with the waves and on non--resonant orbits. If the embedded shearing sheet is initially perturbed by a sinusoidal wave, a swing amplified density wave develops in the disk, which is of the same type as in an isolated sheet or a sheet in a static halo potential, but with an amplitude enhanced by a surprisingly large amount. This appears to be the equivalent of the enhanced bar growth in stellar disks embedded in live dark halos instead of static halo potentials seen in numerical simulations. There is transfer of linear momentum of density waves in the shearing sheet to the halo particles. This is mediated, however, entirely by halo particles on orbits in resonance with the waves similar to the torque exerted by bars on the surrounding halo." }, "0403/astro-ph0403211_arXiv.txt": { "abstract": "We present a multi-wavelength catalog (15 $\\micron$, R, K-band, 1.4 GHz flux) plus spectroscopic identifications for 406 15 $\\micron$ sources detected in the European Large Area {\\it ISO} Survey (ELAIS) region S1, over the flux density range 0.5$<$$S_{15\\micron}$$<$150 mJy. 332 ($\\sim$82\\%) sources are optically identified down to R$\\sim$23.0. Spectra or bona fide stellar identifications are obtained for 290 objects ($\\sim$88\\% of the optically identified sources). The areal coverage, mid-infrared (MIR) and optical completeness of the sample are discussed in order to allow statistical and evolutionary analyses. Two main spectroscopic classes have been found to dominate the MIR extragalactic population: $z<0.5$ star-forming galaxies (from absorbed to extreme starbursts: $\\nu L_{\\nu}(15 \\micron)$$\\approx 10^{8}-10^{11}$~L$_{\\odot}$), which account for $\\sim$75\\% of the sources, and Active Galactic Nuclei (AGN; both type 1 and 2), which account for $\\sim$25\\% of the sources. About 20\\% of the extragalactic sources are dust-enshrouded starburst galaxies [e(a) spectra], and all the starburst galaxies appear more dust extincted in the optical than nearby normal galaxies. We also identified 91 stellar objects ($\\sim$22\\% of the MIR sources). The counts for starburst galaxies and AGN down to 0.6 mJy have been derived. A general trend is found in the optical-MIR spectral energy distribution (SED) of the galaxies, where the MIR-luminous objects have larger MIR to optical luminosity ratios. Based on a variety of analyses, we suggest that the ELAIS sources fainter than R$\\sim$23 are luminous and ultra-luminous MIR galaxies (LIG-ULIGs; $\\nu L_{\\nu}(15 \\micron)$=10$^{11}$-10$^{12}$ L$_\\odot$) at intermediate redshifts ($z$=0.5--1.5), and that consequently the present sample is virtually 100\\% spectroscopically complete up to $z$=0.5. ", "introduction": "The ISOCAM instrument (Cesarsky et al. 1996) on board of the {\\it Infrared Space Observatory} ({\\it ISO}; Kessler et al. 1996) has been able to unveil at 15 $\\micron$ a population of MIR galaxies with two orders of magnitude fainter flux densities than observed by the {\\it Infrared Astronomical Satellite} ({\\it IRAS}). These galaxies result to be ten times more numerous than expected if there were no evolution from $z = 0$ up to $z= 1 - 1.5$ (see Elbaz et al. 1999). These results are confirmed by the detection of a substantial diffuse cosmic Infrared Background in the 140 $\\mu$m - 1 mm range (Puget et al. 1996; Hauser et al. 1998; Fixsen et al. 1998; Lagache et al. 1999), which suggests a strong evolution of the galaxies in the IR (stronger than observed at any other wavelengths). The suggested evolution can either imply a larger fraction of galaxies experiencing an IR luminous phase in the past, or that MIR galaxies were more luminous in the past (see e.g. Chary \\& Elbaz 2001). So far, the nature of the sources responsible for the strong evolution observed in the MIR band has only been studied in small fields, containing sources well known at other wavelengths: the Hubble Deep Field North (HDF-N; Aussel et al. 1999a,b) and South (HDF-S; Oliver et al. 2002; Mann et al. 2002; Franceschini et al. 2003), the Canada-France Redshift Survey (CFRS) 1452+52 field (Flores et al. 1999) and the ELAIS-S2 field (Pozzi et al. 2003). This has yielded a small but meaningful sample of sources sufficient for most multiwavelength studies. Most of these sources can easily be identified with relatively bright optical counterparts ($I < 22.5$) with a median redshift of $z\\simeq0.7-0.8$ imposed by the MIR {\\it k}-correction (Aussel et al. 1999a,b; Flores et al. 1999). However, to obtain reliable source counts and statistically significant information about the nature and evolution of the MIR extragalactic source population, large-area surveys covering a wide range of flux densities are required. In particular, it is very important to bridge the gap between the {\\it IRAS} surveys ($S_{15\\micron}>200$ mJy) and the Deep ISOCAM surveys ($0.1< S_{15\\micron}<2-3$ mJy). ELAIS (Oliver et~al. 2000), with its broad flux range (0.5$$ 2) is {\\it high} (8 out of 18 N $>$ 1 systems at $t=10.5$ Myr) and we therefore test whether our results are compatible with the photometric width of the main sequence in young clusters. The companion frequency varies significantly during the pure $N$-body evolution of the systems. \\item Although we produce a variety of triples, quadruples and higher order multiples, they tend to follow a characteristic pattern of {\\it internal mass distribution}. This involves each binary pair having a mass ratio that does not deviate strongly from unity, and likewise, for quadruples, the mass ratio of the total mass in each binary is not greatly different from unity (i.e. mass ratios in the range $0.5-1$ are favoured in each case). The exception to this is that multiple systems are commonly orbited by {\\it low mass outliers} on eccentric orbits, which are the result of incomplete ejections of low mass objects from multiple systems. We discuss how deep images of the environs of multiple systems, and spectroscopic studies of the primaries of extreme mass ratio binaries, may be used to test this prediction. \\item The multiplicity fraction is an increasing function of primary mass. The brown dwarf desert at {\\it very small separations} is reproduced by our models. \\item We confirm the result of previous simulations that brown dwarf binaries indeed appear to be under-produced by turbulent fragmentation calculations. \\end{itemize} The structure of this paper is as follows. In Section 2 the computational method and initial conditions applied to our models are described. The results on multiple stars are given in Section 3. In Section 4 we perform a detailed comparison of our results with available observational data, and suggest future experiments. Our conclusions are given in Section 5. ", "conclusions": "We have undertaken the first hydrodynamical $+$ $N$-body simulations of multiple star formation that have produced a statistically significant number of stable hierarchical systems, with component separations in the range $\\sim 1-1000$ AU. These simulations have demonstrated that multiple star formation is a major channel for star formation in turbulent flows. The hydrodynamical simulations are followed for $\\approx 0.5$ Myr; subsequently, the remaining gas is removed and the stellar systems followed as $N$-body ensembles for an additional 10 Myr. At this point, all but one of the surviving multiple systems are stable, according to the criteria of Eggleton \\& Kiseleva (1995). We find that the properties of the resulting multiple systems are not significantly sensitive to the large scale geometry of the cloud -- determined by the turbulence -- but rather to the dynamical and competitive accretion processes taking place within the mini-clusters formed out of the collapse and fragmentation of the cloud. At an age of $0.5$ Myr, we find that about $60 \\%$ of stars and brown dwarfs are in multiple systems, with about a third of these being low mass, weakly bound outliers. Excluding these outliers and unbound objects, $7 \\%$ of the remaining objects are in pure binaries ($2$ systems), $14 \\%$ are in quadruples ($2$ systems), $35 \\%$ are in quintuples ($4$ systems), $32 \\%$ are in sextuples ($3$ systems) and $12 \\%$ are in multiples with seven components ($1$ system). The companion frequency is therefore very high, $\\approx 1$. We find that our multiples consist of hierarchies of binaries and triples and that {\\it planetary multiples} (in which companions are not members of binary/triple systems other than the multiple itself) are comparatively rare (occurring $\\sim 25 \\%$ of the time). There is a distinctive pattern of mass distribution within these multiples, with the mass ratio within binaries, and the mass ratios between binaries, rarely deviating far from unity (values of $0.5-1$ are typical). On the other hand, such systems are typically orbited by several low mass outliers (typically at separations of $\\sim 10^4$ AU) on eccentric orbits. About 90\\% of these objects are unstable in timescales of a few $\\times 10^6$ yr (i.e. a few $\\times$ their typical orbital timescale). We find that the $40 \\%$ of objects that are unbound are overwhelmingly of low mass (median mass $\\approx 0.02$ M$_\\odot$). Thus our results imply that in the stellar regime, most stars are in multiples ($\\approx 80\\%$) and that this multiplicity fraction $f_{\\rm m}$ is an increasing function of mass. In this latter respect, these results are qualitatively consistent with a large body of previous works on the decay of small-$N$ systems, both with and without gas (van Albada 1968; McDonald \\& Clarke 1993, 1995; Sterzik \\& Durisen 1998, 2003; DCB03). The high $f_{\\rm m}$ values for GK stars are consistent with adaptive optics measurements of nearby young associations such as MBM~12 and TW~Hydrae (e.g. Brandeker, Jayawardhana \\& Najita 2003), where multiplicity fractions as high 0.64 are found, and radial velocity surveys of visual binaries (e.g. Tokovinin \\& Smekhov 2002) which raise the percentage of spectroscopic sub-systems to at least 40\\%. Low-mass SFR such as Taurus or $\\rho$ Ophiuchus also show companion frequencies in the range $0.3-0.5$, comparable to those predicted by our models at later times. It must be pointed out that the values of the multiplicity fraction $f_{\\rm m}$ for each mass range do not change significantly during the $N$-body evolution of the systems. At an age of 10.5 Myr the fraction of bound and unbound objects has reversed: 40\\% remain in multiples and 60\\% are singles. The companion frequency has dropped to $\\approx 0.3$ due to the ejection of bound outliers to the field. This transference of objects from bound to unbound orbits results in an increase of the number of free floating brown dwarfs by $\\approx 60 \\%$. In this 10 Myr time-span, many multiple systems also experience internal decay: excluding the remaining 3 outliers, 42\\% of the remaining bound objects are in pure binaries (11 systems), 12\\% are in triples (2 systems), 15\\% are in quadruples (2 systems), 19\\% are in quintuples (2 systems) and 12\\% are in sextuples (1 system). We pointed out that low mass stars (and, especially, brown dwarfs) are {\\it locked up} in multiple systems at early times and subsequently released into the field. This remark needs some qualification however. Multiple star formation is hierarchical in our simulations, with structures forming on a particular scale being modified as a result of subsequent merging with structures on a larger scale. It is notable that we find low mass outliers at a separation that is similar to the initial size of the core, and we speculate that if we had modelled a larger volume of cloud, rather than isolated cores, we might have found that these outliers would already have been stripped off by interactions with structures on a larger scale before some of them could settle into stable orbits. This suggests that such outliers should be sought in imaging of relatively isolated pre-main sequence groups and may explain why no such outliers have been detected through deep imaging of multiple systems in Taurus (G. Duch\\^ene, private communication). If such outliers do indeed survive the formation process, then about 10\\% of them are in stable hierarchical orbits at 10.5 Myr. We would thus expect some brown dwarfs to remain in the outer reaches of multiple systems even in the field. Our simulations can thus accommodate the existence of systems with brown dwarfs as wide companions (Gizis et al. 2001) but {\\it only if the primaries of these systems are themselves multiple systems}. (We find that three out of the four bound brown dwarfs present at $t=10.5$ Myr are orbiting a multiple system). We therefore predict that the primaries of binaries containing a brown dwarf in wide orbit should themselves be multiple systems. We have examined how well the products of our simulations compare with the properties of real stellar systems as deduced from the colour magnitude diagram of young clusters (specifically the infrared colour magnitude diagram for Praesepe). Our simulation data compares very favourably with the width of the main sequence in the mass range 0.4-1 M$_\\odot$; indeed the spread of the main sequence in this mass range appears to {\\it require} that stars are commonly assembled into high order multiple systems, although the number of outliers from a pure equal mass binary sequence is not large. The comparison with the Praesepe colour magnitude diagram however illustrates two problems with the simulation results. Firstly, the simulation produces no single star sequence at masses greater than 0.35 M$_\\odot$ (colour bluer than $I-K=2$; there are only two blue circles above 0.35 M$_\\odot$, at 0.581 and 0.584 M$_\\odot$), whereas the observational data shows such a sequence, indicating that single stars and/or low mass ratio binaries are produced in this colour range. We could probably alleviate this problem by modelling a more massive cloud. This would increase the maximum mass of stars produced and enable some stars in the colour range considered to be ejected as singles by encounters with more massive stars. Although this would bring closer agreement with the observed colour magnitude diagram, the lack of low mass ratio binaries in our simulations is in conflict with independent evidence from field binary surveys, such as that of DM91 for G dwarfs. The DM91 data (containing binaries with median separation $30$ AU) showed a distribution with mass ratio ($q$) that rose with decreasing $q$ towards their completeness limit of $q\\sim 0.2$, whereas our mass ratio distribution (with the exception of the very low mass outliers) is strongly concentrated between $0.5$ and unity. This inability to reproduce enough extreme mass ratio systems is a feature of all hydrodynamic modelling of multiple systems to date. It has not however been widely discussed previously, since BBB focused on the close systems, where the observed mass ratio distribution is in any case much more concentrated towards unit mass ratio than for the binary population as a whole (Mazeh et al. 1992; Halbwachs et al. 2003) and where the simulations very naturally reproduce the observed brown dwarf desert at {\\it very small} separations (Marcy \\& Butler 2000; Halbwachs et al. 2000). The second problem revealed by the colour magnitude diagram relates to the fact that the simulations produce almost no binaries with primaries redder than $I-K=2.5$ whereas the data exhibits a pronounced scatter (consistent with large numbers of binaries) in this mass range ($0.15-0.4 M_\\odot$). This scarcity of binaries with low mass primaries in our simulations also conflicts somewhat with the results of binary surveys among M dwarfs (Fischer \\& Marcy 1992; see Figure 4, this paper) and brown dwarfs (Close et al. 2003; Gizis et al. 2003; Mart\\'{\\i}n et al. 2003; Bouy et al. 2003). In summary, we have found that our simulations produce large numbers of hierarchical multiple systems and that relatively isolated young multiples may harbour weakly bound brown dwarf outliers, as a relic of the hierarchical formation process in turbulent flows. We predict that where brown dwarfs are found in wide orbits, the primary should itself turn out to be a multiple. The simulations are consistent with a number of observational constraints: the high (but presently poorly constrained) incidence of hierarchical multiples among field stars and pre-main sequence stars, the absence of brown dwarfs as close companions to normal stars (the brown dwarf desert) and, at a qualitative level at any rate, the positive dependence of the binary fraction on primary mass. There are two areas in which the simulations are not able to replicate observations properly: the simulations under-produce the binary fraction at low masses (M dwarfs and brown dwarfs) and also do not generate enough wide stellar pairs with low mass ratios. ?" }, "0403/astro-ph0403577_arXiv.txt": { "abstract": "{We report on the first simultaneous near-infrared/X-ray detection of the SgrA* counterpart which is associated with the massive 3--4$\\times$10$^6$\\solm ~black hole at the center of the Milky Way. The observations have been carried out using the NACO adaptive optics (AO) instrument at the European Southern Observatory's Very Large Telescope\\footnote{Based on observations at the Very Large Telescope (VLT) of the European Southern Observatory (ESO) on Paranal in Chile; Program: 271.B-5019(A).} and the ACIS-I instrument aboard the \\emph{Chandra X-ray Observatory}. We also report on quasi-simultaneous observations at a wavelength of 3.4~mm using the Berkeley-Illinois-Maryland Association (BIMA) array. A flare was detected in the X-domain with an excess 2 - 8 keV luminosity of about 6$\\times$10$^{33}$~erg/s. A fading flare of Sgr~A* with $>$2 times the interim-quiescent flux was also detected at the beginning of the NIR observations, that overlapped with the fading part of the X-ray flare. Compared to 8-9 hours before the NIR/X-ray flare we detected a marginally significant increase in the millimeter flux density of Sgr A* during measurements about 7-9 hours afterwards. We find that the flaring state can be conveniently explained with a synchrotron self-Compton model involving up-scattered sub-millimeter photons from a compact source component, possibly with modest bulk relativistic motion. The size of that component is assumed to be of the order of a few times the Schwarzschild radius. The overall spectral indices $\\alpha_{NIR/X-ray}$ ($S_{\\nu}$$\\propto$$\\nu^{-\\alpha}$) of both states are quite comparable with a value of $\\sim$1.3. Since the interim-quiescent X-ray emission is spatially extended, the spectral index for the interim-quiescent state is probably only a lower limit for the compact source Sgr~A*. A conservative estimate of the upper limit of the time lag between the ends of the NIR and X-ray flare is of the order of 15~minutes. ", "introduction": "Over the last decades, evidence has been accumulating that most quiet galaxies harbor a massive black hole (MBH) at their centers. Especially in the case of the center of our Galaxy, progress could be made through the investigation of the dynamics of stars (Eckart \\& Genzel 1996, Genzel et al. 1997, 2000, Ghez et al. 1998, 2000, 2003a, 2003b, Eckart et al. 2002, Sch\\\"odel et al. 2002, 2003). Located at a distance of only 8 kpc from the solar system (Reid 1993, Eisenhauer et al. 2003), it allows detailed observations of stars at distances much less than 1~pc from the central black hole candidate, the compact radio source Sgr~A*. Additional strong evidence for a massive black hole at the position of Sgr~A* came from the observation of interim-quiescent and flare activity from that position both in the X-ray and recently in the near-infrared wavelength domain (Baganoff et al. 2001, 2003, Eckart et al. 2003, Porquet et al. 2003, Goldwurm et al. 2003, Genzel et al. 2003, Ghez et al. 2004). Throughout the paper we will use the term 'interim-quiescent' (or IQ) for the apparently constant, low-level flux density states at any given observational epoch since current data cannot exclude flux density variations of that state on longer time scales (days to years). This is especially true for the more compact NIR source (Genzel et al. 2003, Ghez et al. 2004). Simultaneous observations of SgrA* across different wavelength regimes are of high value, since they provide information on the emission mechanisms responsible for the radiation from the immediate vicinity of the central black hole. The first observations of SgrA* covering an X-ray flare simultaneously in the near-infrared using seeing limited exposures revealed only upper limits to the NIR flux density (Eckart et al. 2003). In section 2 of the present paper we report on the first successful simultaneous NIR/X-ray observations using adaptive optics. These observations were for the first time successful in detecting radiation from the SgrA* counterpart both in the NIR and the X-ray wavelength domain. A detailed statistical analysis that supports the simultaneous detection of a SgrA* flare event in the NIR and X-ray domain is given in section 3 of this paper. In section 2 we also describe the quasi-simultaneous mm-observations that were taken just before and after the NIR/X-ray observations. In section 4 we briefly discuss the flux densities and spectral indices we derived from the available data. Section 5 gives a first physical interpretation of the simultaneous detection of SgrA*. A short summary and discussion of the results is given in section 6. ", "conclusions": "We have presented the first successful simultaneous X-ray and NIR detection of SgrA* in a flaring and the IQ-state. The X-ray flare lasted for about 42 minutes and began on 19 June 2003 at about 23:10 UT, more than 4 hours into the X-ray observation. The peak of the flare occurred less than 10 minutes prior to the time of the first VLT/NACO DDT image. The emission during the flare can successfully be described by a SSC model in which the NIR and X-ray flux density excess is produced by up-scattering sub-mm-wavelength photons into the NIR and X-ray domain. With respect to its FWZP duration of 55 - 115 minutes the flare reported here compares well with other flares measured so far. Baganoff et al. (2001), Eckart et al. (2003), and Porquet et al. (2003) report on X-ray events of 45 to 170 minutes. Ghez et al. (2004) and Genzel et al. (2003) report on NIR flare events that last 50 to 80 minutes, respectively. Our newly detected flare event is weaker than most others which have been reported and not necessarily representative of the characteristics of the stronger flares (factor of 50: Baganoff et al. 2001; factor of $>$100: Porquet et al. 2003; Goldwurm et al. 2003). However, during our 2002 \\emph{Chandra} monitoring session we found that flares that are a factor of $>$10 stronger than the quiescent emission occur at a rate of $0.53\\pm0.27$ per day. Weaker flares are more frequent and our newly detected flare event is probably more representative for those weaker flares. Although we have given preference to a simple SSC model in explaining the observed simultaneous flare emission, flux density contributions via other emission mechanisms may be of relevance. Along with enhanced electron heating leading to SSC flares, Markoff et al. (2001) also suggested the possibility that the flares may result from acceleration. This could result from electrons which are energetic enough to account for both the NIR and X-ray flares via direct synchrotron radiation. In fact even SSC models presented by Markoff et al. (2001) and Yuan, Quataert, \\& Narayan (2003) result in a significant amount of direct synchrotron emission in the infrared (see also synchrotron models in Yuan, Quataert, \\& Narayan 2004). The different possible emission mechanisms may also be coupled in a more complicated way. Ambient thermal electrons may be heated during a flare and may produce excess sub-millimeter and infrared flux density. This process could lead to correlated radio/NIR/X-ray variability quite similar to to what is expected in SSC models (Yuan, Quataert, \\& Narayan 2003, 2004, Genzel et al. 2003, Ghez et al. 2004). These synchrotron/ inverse Compton models may also result in small (or no) time lag between the NIR and X-ray emission - compatible with our limit on the time lag for the decaying flank of the flare. In such a scenario it will be difficult to determine the relative importance and flux density contributions of the different emission mechanisms by variability. Since the X-ray source responsible for most of the X-ray IQ-state flux density of SgrA* is extended over a radius of about 0.6~arcsec (Baganoff et al. 2003) its emission can most likely be ascribed to bremsstrahlung from a thermal particle distribution. While the interim-quiescent NIR flux density probably can be attributed to a fairly compact component (Genzel et al. 2003), it is currently not clear how much of the quiescent X-ray flux density originates from that compact region. The extended X-ray component probably originates from hot gas within the $\\sim$1\" Bondi radius that is associated with the accretion flow (see Baganoff et al. 2003 and Quataert 2003 for detailed discussions). If thermal bremsstrahlung is an important mechanism to explain the simultaneous NIR and the X-ray emission, then geometries are conceivable in which the cooler $\\ga$10$^3$~K plasma that gives rise to the NIR emission is spatially offset from the hotter $\\ga$10$^8$~K plasma. One may also speculate about the possibility that the amount of cooler plasma might be larger than the amount of hot plasma. These facts could result in a situation in which the X-ray variations lead the NIR-variations. On the other hand the cooling timescales of the hotter plasma may be larger which would result in the opposite behavior. There are of course also models conceivable that involve small source sizes or apparently cospatial distributions that would result in no measurable time lag or even NIR and X-ray flare events that need not always be correlated (see discussion by Yuan, Quataert, \\& Narayan 2004). If we assume that the (marginally) higher mm-flux density after the flare is related to the flare activity we observed in the X-ray and NIR-domain, then there seems to be a correlation between flare activity in the mm/sub-mm and the X-ray domain in three cases now: the original X-ray flare reported by Baganoff et al. (2001; see also Zhao et al. 2004), the activity around the largest X-ray flare reported by Porquet et al. (2003) and Zhao et al. (2004), and now the small X-ray flare we report on in this publication. However, during an 8-day simultaneous observing campaign using the Owens Valley mm-array at a wavelength of 3-mm and Chandra in the X-ray domain (Mauerhan et al. 2004) did not detect an obvious correspondence of significant fluctuations in both wavelength domains. The currently available data also indicate that the mm-activity may be a function of the X-ray flare magnitude. In the event reported by Porquet et al. (2003) and Zhao et al. (2004) the mm-flux density flared by 100\\% of the continuum following the factor $160$ X-ray flare. Here we report a factor 2 to 3 X-ray flare for which we observed a no more than 10\\% mm-flare following the NIR/X-ray event. Future observations will reveal possible relations between different wavelength regimes. An important question is whether individual mm-flare events are related to events in the NIR or X-ray regime, or whether there are in general more or stronger flares per unit time, when the average mm- or sub-mm flux density is higher. Upcoming simultaneous monitoring programs from the radio to the X-ray regime will be required to further investigate the physical processes that give rise to the observed IQ-state and flare phenomena associated with SgrA* at the position of the massive black hole at the center of the Milky Way." }, "0403/astro-ph0403354.txt": { "abstract": "For a detailed comparison of the appearance of cluster of galaxies in X-rays and in the optical, we have compiled a comprehensive database of X-ray and optical properties of a sample of clusters based on the largest available X-ray and optical surveys: the ROSAT All Sky Survey (RASS) and the Sloan Digital Sky Survey (SDSS). The X-ray galaxy clusters of this RASS-SDSS catalog cover a wide range of masses, from groups of $10^{12.5}$ $M_{\\odot}$ to massive clusters of $10^{15}$ $M_{\\odot}$ in the redshift range from 0.002 to 0.45. The RASS-SDSS sample comprises all the X-ray selected objects already observed by the Sloan Digital Sky Survey (114 clusters). For each system we have uniformly determined the X-ray (luminosity in the ROSAT band, bolometric luminosity, center coordinates) and optical properties (Schechter luminosity function parameters, luminosity, central galaxy density, core, total and half-light radii). For a subsample of 53 clusters we have also compiled the temperature and the iron abundance from the literature. The total optical luminosity can be determined with a typical uncertainty of 20\\% with a result independent of the choice of local or global background subtraction. We searched for parameters which provide the best correlation between the X-ray luminosity and the optical properties and found that the z band luminosity determined within a cluster aperture of 0.5 Mpc $\\rm{h}_{70}^{-1}$ provides the best correlation with a scatter of about 60-70\\%. The scatter decreases to less than 40\\% if the correlation is limited to the bright X-ray clusters. The resulting correlation of $L_X$ and $L_{op}$ in the z and i bands shows a logarithmic slope of 0.38, a value not consistent with the assumption of a constant $M/L$. Consistency is found, however, for an increasing M/L with luminosity as suggested by other observations. We also investigated the correlation between $L_{op}$ and the X-ray temperature obtaining the same result. ", "introduction": "Cluster of galaxies are the largest well defined building blocks of our Universe. They form via gravitational collapse of cosmic matter over a region of several megaparsecs. Cosmic baryons, which represent approximately 10-15\\% of the mass content of the Universe, follow dynamically the dominant dark matter during the collapse. As a result of adiabatic compression and of shocks generated by supersonic motions, a thin hot gas permeates the cluster gravitational potential. For a typical cluster mass of $10^{14}$ $M_{\\odot}$ the intracluster gas reaches a temperature of the order of $10^7$ keV and, thus, radiates optically thin thermal bremsstrahlung and line radiation in the X-ray band. In 1978, the launch of the first X-ray imaging telescope, the \\emph{Einstein} observatory, began a new era of cluster discovery, as clusters proved to be luminous ($\\ge 10^{42-45}$ ergs $\\rm{s}^{-1}$), extended ($\\rm{r}\\sim 1-5$ Mpc) X-ray sources, readily identified in the X-ray sky. Therefore, X-ray observations of galaxy clusters provide an efficient and physically motivated method of identification of these structures. The X-ray selection is more robust against contamination along the line of sight than traditional optical methods since the richest clusters are relatively rare and since X-ray emission, which is proportional to the gas density squared, is far more sensitive to physical overdensities than in the projected number density of galaxies in the sky. In fact the existence of diffuse, very hot X-ray emitting gas implies the existence of a massive confining dark matter halo. Moreover, selection according to X-ray luminosity is also an efficient way to find the highest mass concentrations due to well defined correlation between the X-ray luminosity and the total cluster mass (Reiprich \\& Bh\\\"oringher. 2002). In addition to allowing the identification of galaxy clusters, X-ray observations provide a wealth of information on the intracluster medium itself, e.g. its metal abundance, radial density distribution and temperature profile. These latter quantities, in turn, can be used to reliably estimate the \\emph{total mass} of the system. In addition to the hot, diffuse component, baryons are also concentrated in the individual galaxies within the cluster. These are best studied through photometric and spectroscopic optical surveys, which provide essential information about luminosity, morphology, stellar population and age. Solid observational evidences indicate a strong interaction between the two baryonic components, as galaxies pollute the intracluster medium expelling metals via galactic winds producing the observed metal abundances in clusters (De Grandi et al. 2002, Finoguenov et al. 2001). On the other hand, the evolution of galaxies in clusters is influenced by processes due to the hot gas (e.g. gas stripping by ram pressure, etc.) as it is by internal processes like star formation, galactic winds, supernovae explosions etc., operating inside galaxies themselves (Dressler et al. 1997, Fasano et al. 2000, van Dokkum et al. 2000, Lubin et al. 2002, Kelson et al. 1997,2000, Ziegler\\&Bender 1997, Gomez et al. 2003). In conclusion, understanding the complex physics at play in clusters of galaxies requires combined X-ray and optical observation of a statistically significant sample of these objects. On the basis of these considerations, we have created a large database of clusters of galaxies based on the largest available X-ray and optical surveys: the ROSAT All Sky Survey (RASS), the only X-ray all sky survey ever realized using an imaging X-ray telescope, and the Sloan Digital Sky Survey (SDSS), which is observing the whole Northern Galactic Cap and part of the Southern Galactic Cap in five wide optical bands covering the entire optical range. By carefully combining the data of the two surveys we have created the RASS-SDSS galaxy cluster catalog. Although two galaxy cluster catalogs from the SDSS already exist, the Cut and Enhance Galaxy Cluster Catalog of Goto et al. (2002) and the Merged Cluster Catalog of Bahcall et al. (2003, see also Kim et al. 2002), we prefered to compile a new cluster catalog by selecting the systems in the X-ray band, for which we have reliable X-ray characteristics and for the reasons explained above. The X-ray-selected galaxy clusters cover a wide range of masses, from groups of $10^{12.5}$ $M_{\\odot}$ to massive clusters of $10^{15}$ $M_{\\odot}$ in a redshift range from 0.002 to 0.45. The RASS-SDSS sample comprises all the X-ray detected objects already observed in the sky region covered by the Sloan Digital Sky Survey. One of the first goals is the comparison of the X-ray and the optical appearance of the clusters. We want in particular find optical parameters that provide the closest correlation to the X-ray parameters, such that we can predict within narrow uncertainty limits the X-ray luminosity from these optical parameters and vice versa. So far optical and X-ray cluster surveys have been conducted independently without much intercomparison. Therefore, the empirical relationship between the X-ray luminosity and optical luminosity of clusters is not so well defined, in large part because of the difficulties inherent in measuring the cluster optical luminosity and in getting a homogeneous set of total optical luminosities for a large number of clusters. Edge and Stewart (1991) found that the bolometric X-ray luminosity of a local sample of X-ray-selected clusters correlated very roughly with Abell number and somewhat better with the Bahcall galaxy density (number of bright galaxies within 0.5 $\\rm{h}^{-1}$ Mpc; Bahcall 1977,1981) for the small subsample that had Bahcall galaxy densities. Arnaud et al. (1992) made an heroic effort in computing cluster optical luminosities at low redshift from a heterogenous literature. The first joint X-ray/optical survey of galaxy clusters was the ROSAT Optical X-ray Survey (ROXS, Donahue et al. 2002). They observed 23 ROSAT pointings for a total of 5 square degrees in the I band and partially in the V band. The X-ray selection and the optical selection of cluster candidates were done independently, with the wavelets algorithm in the former case and with a matched filter algorithm in the latter one. They found X-ray and optical coincident detections for 26 galaxy clusters. Donahue et al. (2001) studied the relation between the X-ray luminosity and the matched filter parameter $\\Lambda _{cl}$, which is approximately equivalent to the number of $L^*$ galaxies in the system (Postman et al. 1996). They found a marginally significant correlation between the two quantities with a significant scatter. Yee and Ellingson (2003) defined a new richness parameter as the number of cluster galaxies within some fixed aperture, scaled by a luminosity function and a spatial distribution function. They analysed a sample of 15 clusters from CNOC1 Cluster Redshift Survey, and found a very poor correlation between this parameter and other cluster properties such as the X-ray luminosity, temperature and the velocity dispersion. In the present paper we describe the properties and the information contained in the RASS-SDSS catalog and the resulting correlations between the X-ray and optical properties in the sample. In section 2 we explain how the cluster sample has been created by X-ray selecting the systems from the available X-ray cluster and group catalogs. In section 3 we describe the method for calculating the X-ray cluster properties. We describe in section 4 the optical data and in section 5 the data reduction method. We analyse and discuss the correlations between the optical luminosity and the X-ray properties in section 6. We summarize and discuss the catalog properties and the results in section 7. For all derived quantities, we have used $\\rm{H}_0=70$ $\\rm{km}$$\\rm{s}^{-1}$$\\rm{Mpc}^{-1}$, $\\Omega_ {m}=0.3$ and $\\Omega_ {\\lambda}=0.7$. ", "conclusions": "We created a database of clusters of galaxies based on the largest available X-ray and optical surveys: the ROSAT All Sky Survey (RASS) and the Sloan Digital Sky Survey (SDSS). The RASS-SDSS galaxy cluster catalog is the first catalog which combines X-ray and optical data for a large number (114) of galaxy clusters. The systems are X-ray selected, ranging from groups of $10^{12.5}$ $M_{\\odot}$ to massive clusters of $10^{15}$ $M_{\\odot}$ in a redshift range from 0.002 to 0.45. The X-ray (luminosity in the ROSAT band, bolometric luminosity, redshift, center coordinates) and optical properties (Schechter luminosity function parameters, luminosity, central galaxy density, core, total and half-light radii) are computed in a uniform and accurate way. The catalog contains also temperature and iron abundance for a subsample of 53 clusters from the Asca Cluster Catalog and the Group Sample. The resulting RASS-SDSS galaxy cluster catalog, then, constitutes an important database to study the properties of galaxy clusters and in particular the relation of the galaxy population seen in the optical to the properties of the X-ray luminous ICM. The first investigations reported have shown a tight correlation between the X-ray and optical properties, when the choice of the measurement aperture for the optical luminosity and the optical wavelength band are optimized. We found that the optical luminosity calculated in the i and in the z band correlates better with the X-ray luminosity and the ICM temperature, in comparison to the other Sloan photometric bands. Thus the red optical bands, which are more sensitive to the light of the old stellar population and therefore to the stellar mass of cluster galaxies, have tight correlations with the X-ray properties of the systems. Moreover, we found that the scatter in the $L_{op}-L_X$ and $L_{op}-T_X$ relations can be minimized if the optical luminosity is measured within a cluster aperture between 0.2-0.8 Mpc $\\rm{h}_{70}^{-1}$, with an absolute minimum of the scatter at 0.5 Mpc $\\rm{h}_{70}^{-1}$. The best aperture in the central part of the cluster for the measurement of the optical luminosity is due to the fact that it is a good compromise to simultaneously assess the total richness and the compactness of the cluster. Finally by using the relations obtained from the z band, we demonstrated that, given the optical properties of a cluster, we can predict the X-ray luminosity and temperature with an accuracy of 60\\% and vice versa. By restricting the correlation analysis to the subsample of X-ray detected REFLEX-NORAS clusters, the minimum scatter decreases to less than 40\\% for the $L_{op}-L_X$ relation. Since the observational uncertainties in the optical and in the X-ray luminosity are about 20\\%, the observed scatters in both relations should be intrinsic. The resulting logarithmic slope for the $L_{op}-L_X$ relation with the minimum scatter is $0.38\\pm 0.02$, while the value for the $L_{op}-T_X$ relation is $1.12 \\pm 0.08$. Both results are not consistent with the assumption of hydrostatic equilibrium and a constant M/L. If we assume that M/L depends on the luminosity with the power law $M/L \\propto L^{0.3}$ (Girardi et al. 2002), our results are in very good agreement with the expected values under the assumption of hydrostatic equilibrium. The analysis carried out in this paper on the correlation between X-ray and optical apparence of galaxy clusters is completely empirical. In principle, the best way to proceed in this kind of study is to measure the optical luminosity within the physical size of the cluster, like the virial radius. Without optical spectroscopic data or accurate temperature measurements, the cluster virial radius can be calculated by assuming a theoretical model relating the optical luminosity to the cluster mass. At this stage of the work, we preferred to tackle the cluster X-ray-optical connection with the empirical method explained in the paper, in order to have model-independent results. On the other hand, not taking into account the different cluster sizes could have affected both the slope and the scatter of the given relations (eq. \\ref{eq1}, \\ref{eq2} and \\ref{eq3}). Therefore, for a better understanding of the important connection between the X-ray and optical apparence of galaxy clusters, the optical luminosity has to be calculated within the physical size of the cluster. This work is in progress and will be published in the second paper of this series about the RASS-SDSS galaxy cluster sample. The next step will be the study of the fundamental plane of galaxy clusters. Through this kind of analysis we will find out if the observed scatter in the correlations between the optical and X-ray properties depends on another parameter related to the cluster compactness. Moreover, because of the link between the galaxy cluster fundamental plane and the M/L parameter, we will connect directly the slope of two relations to the behavior of M/L. \\vspace{2cm} Funding for the creation and distribution of the SDSS Archive has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Aeronautics and Space Administration, the National Science Foundation, the U.S. Department of Energy, the Japanese Monbukagakusho, and the Max Planck Society. The SDSS Web site is http://www.sdss.org/. The SDSS is managed by the Astrophysical Research Consortium (ARC) for the Participating Institutions. The Participating Institutions are The University of Chicago, Fermilab, the Institute for Advanced Study, the Japan Participation Group, The Johns Hopkins University, Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, University of Pittsburgh, Princeton University, the United States Naval Observatory, and the University of Washington." }, "0403/astro-ph0403431_arXiv.txt": { "abstract": "Data mining is an important and challenging problem for the efficient analysis of large astronomical databases and will become even more important with the development of the Global Virtual Observatory. In this study, learning vector quantization (LVQ), single-layer perceptron (SLP) and support vector machines (SVM) were put forward for multi-wavelength data classification. A feature selection technique was used to evaluate the significance of the considered features to the results of classification. From the results, we conclude that in the situation of less features, LVQ and SLP show better performance. In contrast, SVM shows better performance when considering more features. The focus of the automatic classification is on the development of efficient feature-based classifier. The classifiers trained by these methods can be used for preselecting AGN candidates. ", "introduction": "Today, there are many impressive archives painstakingly constructed from observations associated with an instrument. The Hubble Space Telescope (HST), the Chandra X-Ray Observatory, the Sloan Digital Sky Survey (SDSS), the Two Micron All Sky Survey (2MASS), and the Digitized Palomar Observatory Sky Survey (DPOSS) are examples of this. Furthermore yearly advances in electronics bring new instruments, doubling the amount of data we collect each year. For example, approximately a gigapixels is deployed on all telescopes today, and new gigapixel instruments are under construction. This trend is bound to continue. Just like what Szalay says, the astronomy is facing \"data avalanche\" (See e.g., Szalay \\& Gray 2001). How to organize, use, and make sense of the enormous amounts of data generated by today's instruments and experiments? It is very time consuming and demands high quality human resources. Therefore, better features and better classifiers are required. In addition, expert systems are also useful to get quantitative information. It is possible to solve the above questions with Neural Networks (NNs), because they permit application of expert knowledge and experience through network training. Furthermore, astronomical object classification based on neural networks requires no priori assumptions or knowledge of the data to be classified as some conventional methods need. Neural networks , over the years, have proven to be a powerful tool capable to extract reliable information and patterns from large amounts of data even in the absence of models describing the data (cf. Bishop 1995) and are finding a wide range of applications also in the astronomical community: catalogue extraction (Andreon et al. 2000), star/galaxy classification (Odewahn et al. 1992; Naim et al. 1995; Miller \\& Coe 1996; M$\\ddot {a}$h$\\ddot {o}$nen \\& Hakala 1995; Bertin \\& Arnout 1996; Bazell \\& Peng 1998), galaxy morphology (Storrie-Lombardi et al. 1992; Lahav et al. 1996), classification of stellar spectra (Bailer-Jones et al. 1998; Allende Prieto et al. 2000; Weaver 2000). Just to name a few, the rising importance of artificial neural networks is confirmed in this kind of task. There is also a very important and promising recent contribution by Andreon et al. (2000) covering a large number of neural algorithms. In this work, a class of supervised neural networks called learning vector quantization (LVQ) was proposed. LVQ shares the same network architecture as the Kohonen self-organizing map (SOM), although it uses a supervised learning algorithm. Bazell \\& Peng (1998) pioneered the use of it in astronomical applications. Another class of supervised neural networks named multi-layer perceptrons (MLP) was presented. Goderya \\& McGuire (2000) summarized progress made in the development of automated galaxy classifiers using neural networks including MLP. Qu et al. (2003) experimented and compared multi-layer perceptrons (MLP), radial basis function (RBF), and support vector machines (SVM) classifiers for solar-flare detection. Meanwhile, an automated algorithm called support vector machines (SVM) for classification was introduced. The approach was originally developed by Vapnik (1995). Wozniak et al. (2001) and Humphreys et al. (2001) have pioneered the use of SVM in astronomy. Wozniak et al. (2001) evaluated SVM, K-means and Autoclass for automated classification of variable stars and compared their effectiveness. Their results suggested a very high efficiency of SVM in isolating a few best defined classes against the rest of the sample, and good accuracy for all classes considered simultaneously. Humphreys et al. (2001) used different classification algorithms including decision trees, K-nearest neighbor and support vector machines for classifying the morphological type of the galaxy. Furthermore, they got the very promising results of their first experiments with different algorithms. Celestial objects radiate energy over an extremely wide range of wavelengths from radio waves to infrared, optical to ultraviolet, X-ray and even gamma rays. Each of these observations carries important information about the nature of objects. Different physical processes show different properties in different bands. Based on these, we apply learning vector quantization (LVQ), single-layer perceptron (SLP) and support vector machines (SVM) to classify AGNs, stars and normal galaxies with data from optical, X-ray, infrared bands. In this paper we present the principles of LVQ, SLP and SVM in section 2. In section 3, we discuss the sample selection and analysis the distribution of parameters. In section 4 the computed results and discussion are given. Finally, in section 5 we conclude this paper with a discussion of general technique and its applicability. ", "conclusions": "Sources classification depends on the quality and amount of real-time data and on the algorithm used to extract generalized mappings. Availability of the high-resolution multi-wavelength data constantly increases. The best possible use of this observational information requires efficient processing and generalization of high-dimensional input data. Moreover, good feature selection techniques, as well as good data mining methods, are in great demand. A very promising algorithm that combines the power of the best nonlinear techniques and tolerance to very high-dimensional data is support vector machines (SVM). In this work we have used histogram as the feature selection technique and applied LVQ, SLP and SVM to multi-wavelength astronomy to classify AGNs from stars and normal galaxies. We conclude that the features selected by histogram are applicable and the performance of SVM models can be comparable to or be superior to that of the NN-based models in the high dimensional space. The advantages of the SVM-based techniques are expected to be much more pronounced in future large multi-wavelength survey, which will incorporate many types of high-dimensional, multi-wavelength input data once real-time availability of this information becomes technologically feasible. All these methods can be used for astronomical object classification, data mining and preselecting AGN candidates for large survey, such as the Large Sky Area Multi-Object Fiber Spectroscopic Telescope (LAMOST). Various data, incuding morphology, photometry, spectral data and so on, can be applied to train the methods and obtain classifiers to classify astronomical objects or preselect intresting objects. When lacking training sets, we may explore some unsupervised methods or outlier finding algorithms to find unusual, rare, or even new types of objects and phenomena. In addition, with the development of the Virtual Observatory, these methods will be part of the toolkits of the International Virtual Observatory." }, "0403/hep-th0403140_arXiv.txt": { "abstract": " ", "introduction": "In AdS/CFT, the $O(2,4)$ isometry group of anti-de Sitter space is reinterpreted as conformal symmetry of the dual theory. For de Sitter space, we do not yet possess the holographically dual theory; nevertheless, we would expect the $O(1,d)$ isometry group of de Sitter space to be a symmetry of that theory. But, unlike AdS, de Sitter space also has horizons, with an associated entropy. There are differing interpretations of the entropy but one view, which we shall explore here, is that it indicates that the dual theory has only a finite number of states\\cite{tombanks,fischler}. However, there is a well-known theorem that says that there are no nontrivial finite-dimensional unitary representations of a noncompact group. Consequently, the Fock space of the dual theory can be either finite-dimensional or a unitary representation of the de Sitter group, but not both. There have thus been claims that de Sitter space has no holographic dual\\cite{gks}, or that the symmetry group is not the de Sitter group\\cite{Witten,alberto}. Here we will consider the possibility that the dual to de Sitter space is indeed defined by a finite-dimensional Fock space, so that the entropy is finite. The finiteness of the Fock space requires that particles in the dual theory obey Fermi-Dirac statistics. Hence we shall construct a spinor dual to de Sitter space. The finitely many Fock states describing a static patch of de Sitter space do not form representations of the de Sitter group but only of the (compact) rotation subgroup. Global de Sitter space is described by two copies of the Fock space. After antipodal identification, these become the space of initial and final states. Some of the de Sitter generators then mix the in- and out-states i.e. bras with kets. The main idea is that there are certain pairings of bras and kets that are nevertheless de Sitter-invariant. These can be used to construct a de Sitter-invariant S-matrix so, in that sense, the isometry group of de Sitter space indeed appears as a symmetry of the dual theory. In this paper, we will illustrate the reconciliation of symmetry and entropy with a toy model based on Dirac spinors; a more refined construction involving fuzzy spheres will appear in a companion paper\\cite{finitereps}. ", "conclusions": "" }, "0403/astro-ph0403695_arXiv.txt": { "abstract": "A primordial cosmological magnetic field induces and supports vorticity or Alfv\\'en waves, which in turn generate cosmic microwave background (CMB) anisotropies. A homogeneous primordial magnetic field with fixed direction induces correlations between the $a_{l-1,m}$ and $a_{l+1,m}$ multipole coefficients of the CMB temperature anisotropy field. We discuss the constraints that can be placed on the strength of such a primordial magnetic field using CMB anisotropy data from the WMAP experiment. We place 3 $\\sigma$ upper limits on the strength of the magnetic field of $B < 15$ nG for vector perturbation spectral index $n=-5$ and $B<1.7$ nG for $n=-7$. ", "introduction": "The origin of the large-scale part of observed galactic magnetic fields, of $\\sim \\mu$G (microGauss) strength and apparently coherent over $\\sim 10$ kpc scales, is unknown. They could be the consequence of nonlinear amplification of a tiny seed field by galactic dynamo processes. An alternate possibility is amplification of a weak seed field through anisotropic protogalactic collapse and subsequent further amplification via galactic differential rotation. In both cases a primordial seed field of strength exceeding $10^{-13}$ to $10^{-12}$ G, coherent over $\\sim $ Mpc scales, is apparently needed, and it is often suggested that upto a $\\sim $ nG strength seed field might be required. See Kulsrud (1999), Widrow (2002), and Giovannini (2003) for reviews of the state of the art in this area. A primordial magnetic field of present strength $\\sim $ nG can leave observable signatures in the cosmic microwave background (CMB) anisotropy. In standard cosmologies vorticity perturbations decay and so do not contribute to CMB temperature or polarization anisotropies. The presence of a cosmological magnetic field generated during an earlier epoch\\footnote{ Quantum fluctuations during an early epoch of inflation can generate a primordial nG magnetic field, coherent over very large scales (see, e.g., Ratra 1992; Bamba \\& Yokoyama 2004)} changes this situation: a primordial magnetic field induces and supports vorticity or Alfv\\'en waves (Adams et al.~1996; Durrer, Kahniashvili, \\& Yates 1998, hereafter DKY). These vector perturbations generate CMB anisotropies.\\footnote{ In the future one can hope to constrain vector modes through their effect on CMB anisotropy polarization anisotropy spectra. CMB polarization spectra that result from vector perturbations due to a primordial magnetic field have been discussed by Seshadri \\& Subramanian (2001), Pogosian, Vachaspati, \\& Winitzki (2002), Mack, Kahniashvili, \\& Kosowsky (2002), and Subramanian, Seshadri, \\& Barrow (2003), while Lewis (2004) considers the case of vector modes supported by free-streaming neutrinos.} The presence of a preferred direction due to a homogeneous magnetic field background leads to an $m$ dependence of $\\langle a_{lm}a_{lm}^*\\rangle$, and induces correlations between the $a_{l+1,m}$ and $a_{l-1,m}$ multipole coefficients of the CMB temperature anisotropy field. Since the CMB anisotropies are observed to be random Gaussian\\footnote{ Colley, Gott, \\& Park (1996), Mukherjee, Hobson, \\& Lasenby (2000), and Park et al.~(2001) are some early discussions of the Gaussianity of the CMB anisotropy. More recent discussions of the Gaussianity of the WMAP CMB anisotropy data are in Komatsu et al.~(2003), Colley \\& Gott (2003), Chiang et al.~(2003), Park (2004), Eriksen et al.~(2004a, 2004b), Coles et al.~(2004), Vielva et al.~(2004), Copi, Huterer, \\& Starkman (2003), Hansen et al.~(2004), Gurzadyan et al.~(2004), and Mukherjee \\& Wang (2004). The simplest inflation models predict Gaussian fluctuations (see, e.g., Fischler, Ratra, \\& Susskind 1985; Ratra 1985), and this is consistent with most observational indications (see, e.g., Peebles \\& Ratra 2003). While there are indications of mild peculiarities in some subsets of the WMAP data, for example the apparent paucity of large-scale power (Spergel et al.~2003; see G\\'orski et al.~1998a for a similar indication from COBE data) and the differences between data from different parts of the sky (see papers cited above), foreground contamination (see, e.g., Park, Park, \\& Ratra 2002; Mukherjee et al.~2002, 2003; de Oliveira-Costa et al.~2003; Bennett et al.~2003b; Tegmark, de Oliveira-Costa, \\& Hamilton 2003) and other systematics might be responsible for part of this.}, it is known that such a contribution can only be subdominant. We use the observed $\\overline{\\langle a_{l-1,m}a_{l+1,m}^* \\rangle}$ correlations measured by WMAP to place constraints on the strength of a homogeneous primordial magnetic field. The angular brackets here denote an ensemble average, and the overbar indicates an average over $m$ for each $l$. Limited by cosmic variance uncertainties, this would be the useful measure to characterize the signature of a homogeneous primordial magnetic field. The model on which we base our analysis is introduced in \\S 2. Our analysis of the WMAP data and our results are discussed in \\S 3. We conclude in \\S 4. ", "conclusions": "We study off-diagonal correlations of the form $D_l=\\overline{\\langle a_{l-1,m}a_{l+1,m}^* \\rangle}$ in the first year WMAP CMB anisotropy data. Such correlations can result from a homogeneous primordial magnetic field. We do not find significant off-diagonal correlations in the data, which appear to be satisfactorily fit by a zero primordial magnetic field hypothesis. We place 3 $\\sigma$ upper limits on the strength of the magnetic field of $B < 15$ nG for spectral index $n=-5$ and $B<1.7$ nG for $n=-7$. These two cases are interesting as they correspond to a Harrison-Peebles-Yu-Zel'dovich scale-invariant spectrum result for the $C_l$'s and $D_l$'s, and to a possible inflation model primordial magnetic field perturbation spectrum, respectively. These two cases also span the range of constraints that can be placed on $B$ using this method. Future CMB anisotropy data should allow for tighter constraints on a primordial cosmological magnetic field. \\bigskip We acknowledge useful discussions with R.~Durrer and A.~Kosowsky. GC, TK, and BR acknowledge support from NSF CAREER grant AST-9875031 and DOE EPSCoR grant DE-FG02-00ER45824. TK also acknowledges CRDF-GRDF grant 3316. PM and YW acknowledge support from NSF CAREER grant AST-0094335." }, "0403/astro-ph0403140_arXiv.txt": { "abstract": "We have followed the evolution of multi-mass star clusters containing massive central black holes by N-body simulations on the GRAPE6 computers of Tokyo University. We find a strong cluster expansion and significant structural changes of the clusters. Clusters with IMBHs have power-law density profiles $\\rho \\sim r^{-\\alpha}$ with slopes $\\alpha=1.55$ inside the influence sphere of the central black hole. This leads to a constant density profile of bright stars in projection, which rules out the presence of intermediate mass black holes in core collapse clusters. If the star clusters are surrounded by a tidal field, a central IMBH speeds up the destruction of the cluster until a remnant of a few hundred stars remains, which stays bound to the IMBH for a long time. We also discuss the efficiency of different detection mechanisms for finding IMBHs in star clusters. ", "introduction": "X-ray observations of starburst and interacting galaxies have revealed a class of ultra-luminous X-ray sources (ULX), with luminosities of order $L\\approx 10^{39}$ to $10^{41}$ \\citep{Makishimaetal2000}. If the flux is radiated isotropically, this exceeds the Eddington luminosities of stellar mass black holes by orders of magnitude, making ULX good candidates for IMBHs. Many ULX appear to be associated with star clusters \\citep{Fabbianoetal1997}, the irregular galaxy M82 for example hosts an ULX with luminosity $L > 10^{40}$ erg/sec near its center \\citep{Matsumotoetal2001, Kaaretetal2001} whose position coincides with that of the young ($T \\approx 10$ Myrs) star cluster MGG-11. \\citet{PortegiesZwartetal2004a} and \\citet{McMillanetal2004} have performed $N$-body simulations of several star clusters in M82 and showed that runaway merging of massive stars could have led to the formation of an IMBH with a few hundred to a few thousand solar masses in MGG-11, thereby explaining the presence of the ultraluminous X-ray source. The fact that a considerable fraction of star cluster might have formed intermediate mass black holes (IMBHs) has interesting consequences. For example, IMBHs of a few 100 to a few 1000 $M_\\odot$ would explain why the mass-to-light ratios in several globular clusters increase towards the center \\citep{Gerssenetal2002, Colpietal2003}, although the data presented so far is also compatible with an unseen concentration of neutron stars and heavy mass white dwarfs \\citep{Baumgardtetal2003}. IMBHs in star clusters would also be prime targets of the forthcoming generation of ground and space-based gravitational wave detectors and could provide the missing link between the stellar mass black holes formed as the end product of stellar evolution and the $10^6$ to $10^9 M_\\odot$ sized black holes found in galactic centers \\citet{Ebisuzakietal2001}. In this paper we explore the dynamical evolution of star clusters containing massive black holes. We study how star clusters evolve during a Hubble time and compare the outcome of our simulations with galactic globular clusters in order to determine which clusters are likely to contain IMBHs. We also study what is left bound to an IMBH after the parent cluster is dissolved and discuss ways how to detect an IMBH in a globular cluster. ", "conclusions": "We have performed two sets of $N$-body simulations of multi-mass star clusters containing intermediate mass black holes. We found that the 3-dimensional mass-density follows a $\\rho \\sim r^{-1.55}$ density profile around the central black hole. When viewed in projection, the luminosity profile of clusters with massive black holes has a constant density core. The presence of intermediate mass black holes in core collapse globular clusters like M15 is therefore ruled out by our simulations. As was shown in \\citet{Baumgardtetal2003}, a more natural explanation for mass-to-light ratios that increase towards the center in such clusters is a dense concentration of neutron stars, white dwarfs and stellar mass black holes. The detection of a central black hole through proper motion or radial velocity measurements of stars in the central cusp around the black hole is possible with HST for the nearest globular clusters. It might also be possible to find black holes in globular clusters by their gravitational wave emission. The detection through the X-ray emission arising from the IMBH is possible only after the tidal disruption of a star or when a star captured through tidal heating is in a close enough orbit to the IMBH. Intermediate mass black holes also speed up the dissolution of star clusters if the clusters are surrounded by a tidal field." }, "0403/astro-ph0403230_arXiv.txt": { "abstract": "An Artificial Neural Network (ANN) scheme has been employed that uses a supervised back-propagation algorithm to classify 2000 bright sources from the Calgary database of IRAS (Infrared Astronomical Satellite) spectra in the region $8\\mu$m to $23\\mu$m. The database has been classified into 17 predefined classes based on the spectral morphology. We have been able to classify over 80 percent of the sources correctly in the first instance. The speed and robustness of the scheme will allow us to classify the whole of the LRS database, containing more that 50,000 sources, in the near future. ", "introduction": "Infrared Astronomical Satellite Low Resolution Spectrometer (LRS) recorded spectra of some 50,000 sources in blue ($8 - 15\\mu$m) with $\\lambda/\\Delta\\,\\lambda \\sim 40$ and in red ($13 - 23\\mu$m) with a resolution $\\sim 20$. A total of 5425 objects with better quality spectra were included in the Atlas of low-resolution IRAS spectra (1986, hereafter the Atlas). Volk \\& Cohen (1989a) published spectra of 356 IRAS point sources with $\\rm F_{\\nu} (12 \\mu m) > 40 Jy$ that were not included in the Atlas. These brighter sources were classified into nine classes based upon the spectral morphology. Sixty percent of the sources have silicate emission and red-continuum spectra associated with H II region sources. No emission-line sources formed part of the set of 356 spectra. This sample was also used to test the classification scheme of IRAS sources based on broad-band colors. Classifiable spectra were found for 338 of the sources in the sample. The remaining 18 sources had either extremely noisy or incomplete spectra. It was found that some class of sources overlapped on the color-color diagrams and, therefore, the nature of some of the IRAS sources could not be determined from the IRAS photometry. Volk et al. (1991) published an additional 486 spectra belonging to sources with $\\rm 12 \\mu m$ fluxes between $20$ and $\\rm 40 Jy$ that were also not in the Atlas. Classifiable spectra were found for 424 sources. The spectra were classified into nine groups as in Volk \\& Cohen (1989a) that describe the astrophysical nature of these sources. Kwok, Volk \\& Bidelman (1997) processed 11,224 spectra (including sources in the Atlas), corresponding to a flux limit of $\\rm 7 Jy$ at $\\rm 12 \\mu m$. These spectra were also classified by-eye and put into nine classes based on the presence of emission and absorption features and on the shape of the continuum. They identified optical counterparts of these IRAS sources in the existing optical and infrared catalogs and listed the optical spectral types if they were known. It is evident that large databases like the one referred to above require automated schemes for any analysis. Artificial Neural Networks (ANN) have been employed extensively in several branches of Astronomy for automated data analysis (Lahav \\& Storrie-Lombardi, 1994). ANN have been used previously by the IUCAA group in three distinct areas of stellar Astronomy. They have been applied to classify digitized optical and ultraviolet spectra (Gulati et. al. 1994a,b; Singh et al. 1998); to compare a set of observed spectra of F \\& G dwarfs with a library of synthetic spectra (Gulati et al. 1997a, Gupta et al. 2001); and to determine the reddening properties of hot stars from the low-dispersion ultraviolet spectra (Gulati et al. 1997b). We have attempted to classify 2000 brightest sources from the Atlas into 17 classes by means of Artificial Neural Networks. In the next section we describe features of the 17 new classes. In Section 3, we present details of the ANN scheme. Results are discussed in Section 4 and important conclusions of the study are presented in the last section. Further, an appendix has been added just before the reference section, to explain the general ANN architecture for the benefit of the readers. ", "conclusions": "We have demonstrated in this paper the application of ANN scheme to a large database of 2000 IRAS spectra and have been able to correctly classify more than 80\\% of the data-set. The misclassified spectra were looked in detail and most of them could have been wrongly catalogued or had features which would have been confusing for even a human classifier. We stress here that the speed and robustness of this scheme can be very useful for classifying the whole of LRS database containing over 50,000 sources." }, "0403/gr-qc0403088_arXiv.txt": { "abstract": "Japanese laser interferometric gravitational wave detectors, TAMA300 and LISM, performed a coincident observation during 2001. We perform a coincidence analysis to search for inspiraling compact binaries. The length of data used for the coincidence analysis is 275 hours when both TAMA300 and LISM detectors are operated simultaneously. TAMA300 and LISM data are analyzed by matched filtering, and candidates for gravitational wave events are obtained. If there is a true gravitational wave signal, it should appear in both data of detectors with consistent waveforms characterized by masses of stars, amplitude of the signal, the coalescence time and so on. We introduce a set of coincidence conditions of the parameters, and search for coincident events. This procedure reduces the number of fake events considerably, by a factor $\\sim 10^{-4}$ compared with the number of fake events in single detector analysis. We find that the number of events after imposing the coincidence conditions is consistent with the number of accidental coincidences produced purely by noise. We thus find no evidence of gravitational wave signals. We obtain an upper limit of 0.046 /hours (CL $= 90 \\%$) to the Galactic event rate within 1kpc from the Earth. The method used in this paper can be applied straightforwardly to the case of coincidence observations with more than two detectors with arbitrary arm directions. ", "introduction": "In the past several years, there has been substantial progress in gravitational wave detection experiments by the ground-based laser interferometers, LIGO\\cite{ref:LIGOdescription}, VIRGO\\cite{ref:virgo}, GEO600\\cite{ref:geo}, and TAMA300\\cite{ref:tama, ref:ando}. The observation of gravitational waves will not only be a powerful tool to test general relativity, but also be a new tool to investigate various unsolved astronomical problems and to find new objects which were not seen by other observational methods. The Japanese two laser interferometers, TAMA300 and LISM, performed a coincident observation during August 1 and September 20, 2001 (JST). Both detectors showed sufficient stability that was acceptable for an analysis to search for gravitational wave signals. Given the sufficient amount of data, it was a very good opportunity to perform a coincidence analysis with real interferometers' data. There were several works to search for gravitational waves using interferometeric data. A coincidence analysis searching for generic gravitational wave bursts in a pair of laser interferometers has been reported in \\cite{ref:Nicholson}. Allen et al.\\cite{ref:40m} analyzed LIGO 40m data and obtained an upper limit of 0.5/hour (CL = 90$\\%$) on the Galactic event rate of the coalescence of neutron star binaries with mass between 1$M_\\odot$ and 3$M_\\odot$. Tagoshi et al.\\cite{ref:DT2} analyzed TAMA300 data taken during 1999 and obtained an upper limit of 0.59/hour (CL = 90$\\%$) on the event rate of inspirals of compact binaries with mass between 0.3$M_\\odot$ and 10$M_\\odot$ and with signal-to-noise ratio greater than 7.2. Very recently, an analysis using the first scientific data of the three LIGO detectors was reported~\\cite{ref:LIGOinspiral}, and an upper limit of $1.7\\times 10^2$ per year per Milky Way Equivalent Galaxy is reported. Recently, International Gravitational Event Collaboration (IGEC) of bar detectors reported their analysis using four years of data to search for gravitational wave bursts \\cite{ref:IGEC03}. They found that the event rate they obtained was consistent with the background of the detectors' noise. In the matched filtering analysis using real data of single laser interferometer (e.g. \\cite{ref:40m}, \\cite{ref:DT2}), many fake events were produced by non-Gaussian and non-stationary noise. In order to remove such fake events, it is useful to perform coincidence analysis between two or more independent detectors. Furthermore, coincidence analysis is indispensable to confirm the detection of gravitational waves when candidates for real gravitational wave signals are obtained. The purpose of this paper is to perform coincidence analysis using the real data of TAMA300 and LISM. We consider gravitational waves from inspiraling compact binaries, comprized of neutron stars or black holes. They are consider to be one of the most promising sources for ground based laser interferometers. Since the waveforms of the inspiraling compact binaries are known accurately, we employ the matched filtering by using the theoretical waveforms as templates. Matched filtering is the optimal detection strategy in the case of stationary and Gaussian noise of detector. However, since the detectors' noise is not stationary and Gaussian in the real laser interferomters, we introduce $\\chi^2$ selection method to the matched filtering. We analyze the data from each detector by matched filtering which produces event lists. Each event is characterized by the time of coalescence, masses of the two stars, and the amplitude of the signal. If there is a real gravitational wave event, there must be an event in each of the event list with consistent values of parameters. We define a set of coincidence conditions to search for coincident events in the two detectors. We find that we can reduce the number of events to about $10^{-4}$ times the original number. The coincidence conditions are tested by injecting the simulated inspiraling waves into the data and by checking the detection efficiency. We find that the detection efficiency is not affected significantly by imposing the coincidence conditions. We estimate the number of coincident events produced accidentally by the instrumental noise. By using a technique of shifting the time series of data artificially, we find that the number of events survived after imposing the coincidence conditions is consistent with the number of accidental coincidences produced purely by noise. We propose a method to set an upper limit to the real event rate using results of the coincidence analysis. In the case of TAMA300 and LISM, we obtain an upper limit of the event rate as 0.046/hour (CL = 90$\\%$) for inspiraling compact binaries with mass between 1$M_\\odot$ and 2$M_\\odot$ which are located within 1kpc from the Earth. In this case, since TAMA300 is much more sensitive than LISM, the upper limit obtained from the coincidence analysis is less stringent than that obtained from the TAMA300 single detector data analysis. This is because the detection efficiency in the coincidence analysis is determined by the sensitivity of LISM. Thus, the upper limit obtained here is not the optimal one which we could obtain using the TAMA300 data taken during 2001. The method to set an upper limit to the event rate proposed here can be extended straightforwardly to the case of a coincidence analysis for a network of interferometric gravitational wave detectors. This paper is organized as follows. In Section~\\ref{sec:detector}, we briefly describe the TAMA300 and LISM detectors. In Section~\\ref{sec:method}, we discuss a method of matched filtering search used for TAMA300 and LISM data. In Section~\\ref{sec:matchedfilterresults}, the results of the matched filtering search for each detector are shown. In section~\\ref{sec:coincidence}, we discuss a method of the coincidence analysis using the results of single-detector searches, and the result of the coincidence analysis is shown. We also derive the upper limit to the event rate in Section~\\ref{sec:limit}. Section~\\ref{sec:summary} is devoted to summary. In Appendix A, we discuss a $\\chi^2$ veto method to distinguish between real events and fake events produced by non-Gaussian noise. In Appendix B, we examine a different choice of $\\Delta t$ ( the length of duration to find local maximum of matched filtering output ) for comparison. In Appendix C, we discuss a sidereal time distribution of coincidence events. In Appendix D, we review a method to estimate the errors in the parameters due to noise using the Fisher matrix. Throughout this paper, the Fourier transform of a function $h(t)$ is denoted by $\\tilde{h} (f)$, which is defined by \\begin{equation} \\tilde{h} (f) = \\int_{-\\infty}^{\\infty} dt\\ e^{2 \\pi i f t} h(t) . \\end{equation} ", "conclusions": "\\end{table} \\begin{table}[htpd] \\begin{center} \\renewcommand{\\arraystretch}{1.5} \\begin{tabular}{c c c c c} \\hline \\hline & Year & period & obsevation time [hours] &Topics \\\\ \\hline DT1&1999& 6-7 Aug. & 11 &Total detector system check \\\\ &&&&and Calibration test\\\\ DT2&1999& 17-20 Sept.&31& First event search \\\\ DT3&2000& 20-23 April&13& Sensitivity improved\\\\ DT4&2000& 21 Aug.-4 Sept.&167 &100 hours observation\\\\ DT5&2001& 2-10 Mar. &111& Full time observation\\\\ DT6&2001& 1 Aug.-20 Sept.&1038& 1000 hours observation \\\\ &&&&and coincident observation with LISM\\\\ DT7&2002& 31 Aug.-2 Sept.&25& Power recycling installed (Full configuration)\\\\ DT8&2003& 14 Feb.-14 April & 1158 & Coincident observation with LIGO \\\\ DT9&2003 - 2004& 28 Nov.- 10 Jan. & 557 & Full automatic operation\\\\ & & & &and Partial coincident observation \\\\ &&&&with LIGO and GEO600 \\\\ \\hline \\hline \\end{tabular} \\end{center} \\caption{Observation history of TAMA300} \\label{tab:history} \\end{table} \\subsection{LISM}\\label{sec:detectorl} LISM is a laser interferometer gravitational wave antenna with arm length of 20m, located in the Kamioka mine ($36.25^\\circ$N, $137.18^\\circ$E), 219.02km west of Tokyo. The detector's arm orientation is $165^\\circ$ measured counter clockwise from East. The LISM antenna was originally developed as a prototype detector from 1991 to 1998 at the National Astronomical Observatory of Japan, in Mitaka, Tokyo, to demonstrate advanced technologies \\cite{ref:sato}. In 1999, it was moved to the Kamioka mine in order to perform long-term, stable observations. Details of the LISM detector is found in \\cite{ref:satoLISM}. The laboratory site is 1000m underground in the Kamioka mine. The primary benefit of this location is extremely low seismic noise level except artificial seismic excitations. Furthermore, much smaller environmental variations at this underground site are beneficial to stable operation of a high-sensitivity laser interferometer. The optical configuration is the Locked Fabry-Perot interferometer. The finesse of each arm cavity was about 25000 to have a cavity pole frequency of 150Hz. The main interferometer was illuminated by a Nd:YAG laser yielding 700mW of output power, and the detector sensitivity spectrum was shot-noise limited at frequencies above about 1kHz. The operation of LISM was started in early 2000, and has repeatedly been tested and improved since. The data used in this analysis were taken in the observations between August 1st and 23th and between September 3rd and 17th, 2001. The total length of data is 780 hours. The first half of the period was in a test-run and some improvements were made after that. The data from the second half were of good quality to be suitable for a gravitational wave event search, so 323 hours of data for the latter half was dedicated for this analysis. The best sensitivity during this period was about $h\\sim 6.5 \\times 10^{-20}/\\sqrt{\\rm{Hz}}$ around 800Hz. \\label{sec:summary} In this paper, we performed a coincidence analysis using the data of TAMA300 and LISM taken during DT6 observation in 2001. We analyzed the data from each detector by matched filtering and obtained event lists. Each event in the lists was characterized by the time of coalescence, masses of the two stars, and the amplitude of events. If any of the events are true gravitational wave events, they should have the consistent values of these parameters in the both event lists. We proposed a method to set coincidence conditions for the source parameters such like the time of coalescence, chirp mass, reduced mass, and the amplitude of events. We took account of the time delay due to the distance between the two detectors, the finite mesh size of the mass parameter space, the difference in the signal amplitudes due to the different sensitivities and antenna patterns of the detectors, and errors in the estimated parameters due to the instrumental noise. Our Monte Carlo studies showed that we would not lose events significantly by imposing the coincidence conditions. By applying the above method of the coincidence analysis to the event lists of TAMA300 and LISM, we can reduce the number of fake events by a factor $10^{-4}$ compared with the number of fake events before the coincidence analysis. In order to estimate the number of accidental coincidences produced by noise, we used the time shift procedure. We found that the number of events survived after imposing the coincidence conditions is consistent with the expected number of accidental coincidences within the statistical fluctuations. Thus we found no evidence of gravitational wave signals. As discussed in Appendix~\\ref{sec:sidereal}, the sidereal time distribution of the survived events were also consistent with the distribution of accidentals. Finally, we proposed a simple method to set an upper limit to the event rate and applied it to the above results of the coincidence analysis. We obtained an upper limit to the Galactic event rate within 1kpc from the Earth to be 0.046 [1/hour] (90\\%\\,CL). In our case, since LISM has a much lower sensitivity than TAMA300, we were unable to obtain a more stringent upper limit to the event rate than the one obtained by the single-detector analysis of TAMA300. This is because the detection efficiency in the coincidence analysis is determined by the detector with a lower sensitivity. However, if we have two detectors that have comparable sensitivities, it is possible to obtain an improved upper limit compared to a single-detector analysis. As an example, let us imagine the case when the sensitivity of LISM is the same as that of TAMA300. The result of Galactic event simulations suggests that the detection efficiency in the case of a single-detector analysis is 0.35, while it improves to 0.48 in the case of a coincidence analysis. These values are translated to upper limits on the Galactic event rate of 0.026 [1/hour] ($90\\%$\\,CL) for the single-detector case and 0.019 [1/hour] ($90\\%$\\,CL) for the two-detector case. The method of a coincidence analysis and the method to set an upper limit to the event rate proposed here can be readily applied to the case when there are more than two detectors with arbitrary arm directions. Hence these methods will be useful for data analysis for a network of interferometeric gravitational wave detectors in the near future." }, "0403/astro-ph0403006_arXiv.txt": { "abstract": "Gauge invariant quintessence perturbations in the quintessence and cold dark matter ($QCDM$) model are investigated. For three cases of constant equation-of-state (EOS) parameter, linear scalar field potential, and supergravity scalar field potential, their perturbation evolutions have a similar dependence on EOS parameter and scale, but they have different sensitivity to the initial conditions due to the different shapes of the quintessence potential. They have a minor effect on primary CMB anisotropies, but change the secondary CMB effect in different ways. The first case is insensitive to initial quintessence perturbations and only modifies the Integrated Sachs-Wolfe (ISW) effect within a factor of 2. The other two cases are sensitive to initial conditions at large scales and could affect the secondary CMB anisotropies drastically depending on how smooth the initial perturbations are. This makes it possible for future cosmological probes to provide constraints on quintessence properties. ", "introduction": "Recent observations of an accelerating universe \\cite{Riess1998, Perlmutter1999, Riess2001} imply the existence of dark energy characterized by a negative pressure to density ratio, also known as the EOS parameter $w$. There have been discussions of the cosmological constant, the dark energy evolving according to a specific EOS parameter \\cite{Silveira1997, Chiba1997}, and the dark energy consisting of a dynamical cosmic scalar field, the quintessence \\cite{Ratra1988, Ferreira1997}. In contrast to the cosmological constant, the quintessence component in $QCDM$ model has kinematic behavior and can develop perturbations. It has important features of time-evoluting EOS parameter and scale-dependent effective sound speed \\cite{Caldwell1998a, Caldwell1998b}. One significant issue is the sensitivity to initial perturbation conditions, which attracts much attention. For some $QCDM$ models, the quintessence potentials can be approximated with constant EOS parameter, and the perturbations evolutions are insensitive to the initial quintessence perturbations, with minor effect on observations \\cite{Dave2002, DaveThesis, Weller2003}. For exponential potential $V(Q)=\\hat{V} e^{-(c/M)Q}$ \\cite{Wetterich1985, Wetterich1988, Ratra1988}, although quintessence perturbations may stay non-zero at some time for different background attractor solutions, they die out in matter-dominant time for physically reasonable models \\cite{Tassilo2001}, which shows the insensitivity again. For the tracking quintessence models \\cite{Brax2000}, large-scale non-adiabatic perturbations can grow or stay constant before entering tracker regime, but have to get suppressed afterwards \\cite{MalquartiLiddle2002}. However, the first-order matrix formulation developed latter \\cite{Bartolo2003} applies to general quintessence potential cases on all linear scales and indicates the possibility of non-vanishing entropy perturbations. Recently some other $QCDM$ models with special quintessence potentials, such as linear $V(Q)$ \\cite{Dimopoulos2003} and supergravity $V(Q)$ \\cite{Kallosh2002}, lead to special background evolutions. So it will be interesting to see the perturbations behavior and their sensitivity to initial conditions in these models, as well as their effects on cosmic microwave background (CMB) \\cite{Bennett2003} and baryon mass power spectrum \\cite{Scranton2003} in the universe. In this paper, we describe the universe by a simple $QCDM$ model with the conformal-Newtonian metric in section 2. The cosmological perturbations are solved for constant $w$ cases in section 3, for linear-$V(Q)$ cases in section 4, and for supergravity-$V(Q)$ cases in section 5. The results are analyzed by the matrix formulation and their effects on cosmological observations are discussed in section 6. ", "conclusions": "In another viewpoint, we can apply the first-order matrix formulation in \\cite{Bartolo2003} to analyze the different quintessence perturbation behaviors in above $QCDM$ models. At large scales ($k \\ll 1$), the relative entropy perturbation $S$ and intrinsic entropy perturbation $\\Gamma$ \\cite{Wands2000, Malik2003, Kodama1984} have coupled evolution equation $$ \\frac{\\partial}{\\partial \\ln(a/a_0)} \\left( \\begin{array}{c} S \\\\ \\Gamma \\end{array} \\right) = 3 \\left( \\begin{array}{cc} w_Q - w_f + \\gamma_f \\Omega_f (w_f - c_{sQ}^2) / \\gamma & \\gamma_f \\Omega_f (1 - c_{sQ}^2) / \\gamma \\\\ - \\gamma /2 & w_Q - \\gamma /2 \\end{array} \\right) \\times \\left( \\begin{array}{c} S \\\\ \\Gamma \\end{array} \\right) $$ where $_f$ means the perfect fluid component, $\\gamma_i \\equiv 1+w_i$, and $c_{sQ}^2 \\equiv \\dot{p} / \\dot{\\rho}$. For most of the time in our $QCDM$ models, $CDM$ dominates and have $w_f \\simeq 0$, $\\Omega_f \\simeq \\gamma_f \\simeq \\gamma \\simeq 1$, and $-13\\sigma$ significance level from high- to low-density environments. We explain this trend as the combination of two related effects: a manifestation of the morphology-density relation whereby the fraction of early-type galaxies which have shallow faint-end slopes increases with density, at the expense of late-type galaxies which have steep faint-end slopes; and a luminosity-segregation due to dwarf galaxies being cannibalised and disrupted by the cD galaxy and the ICM in the cluster core \\cite{lopezcruz}. To separate the two effects, we consider the TSLF of galaxies belonging to the cluster red sequence, as these are predominately early-type galaxies, and so any trends with environment should be independent of the morphology-density relation. Some luminosity-segregation is observed, with a reduction in the fraction of dwarf galaxies in the high-density regions, as manifested by their shallower TSLF faint-end slopes. This effect is smaller than that for the overall LF, being significant only at the $1.7\\sigma$ level, indicating that both luminosity-segregation and the morphology-density relation drive the observed trends in the overall LF. Luminosity-segregation is predicted by simulations for early-type galaxies in clusters, with bright early-type galaxies much more concentrated than their faint counterparts which follow the distribution of the dark matter mass profile (Springel \\etal \\cite{springel}). In a study of 45 low-redshift \\mbox{$(0.04$1300 known pulsars in the Galaxy are Vela-like \\citep[i.e. with $P \\sim 100$\\,ms, $\\dot{E} \\sim 10^{36}-10^{37}$ ergs s$^{-1}$, and characteristic ages $\\tau_c \\sim$10$-$20\\,kyr, see][]{kbm+03}. Half of these are known to have associated PWN \\citep*{krh04}. Here we present the results of X-ray observations made with the {\\it Chandra} ACIS-S detector. These observations resolve the {\\it ASCA} source into a new PWN, which we name \\pwn , and a thermally emitting point source, likely the neutron-star surface. {\\it Chandra} observations in continuous-clocking mode show a potential detection of X-ray pulsations from \\psr . We also present radio timing observations made with the Arecibo telescope, which show that \\psr\\ has glitched and is highly scattered. ", "conclusions": "We have used the {\\it Chandra X-ray Observatory} to observe the young and energetic pulsar \\psr\\ and have detected a new PWN, \\pwn , as well as an embedded point source near the radio timing position of the pulsar. The morphology of \\pwn\\ is perhaps best decribed as an equatorial torus, and we have derived geometrical parameters for the orientation of the system by spatially fitting the nebula. Spectral fitting of the nebula shows that it is similar to the nebulae of other energetically-similar pulsars. Spectral fitting of the point source shows that it is well fit by an absorbed black-body model or a black-body plus power-law model. Radio timing observations of \\psr\\ have revealed a glitch similar to the largest glitches seen in the Vela pulsar. Analysis of countinuous-clocking data from {\\it Chandra} indicates that there may be weak X-ray pulsations from the source (pulsed fraction of $\\sim$37\\%) at the period predicted by the radio ephemeris. \\psr\\ remains the most likely counterpart to the high-energy \\EG\\ \\gr\\ source GeV J2020+3651 (which is the 10th brightest \\gr\\ source above 1\\,GeV). It is an excellent \\gr\\ pulsar candidate for the AGILE and GLAST missions, although a contemporaneous ephemeris will be essential, given the large amount of timing noise the pulsar exhibits. Deeper observations with {\\it Chandra} and {\\it XMM-Newton} are needed to clarify the morphology of \\pwn\\ and to confirm the presence of pulsations. Such observations are also able to confirm whether the faint outer jet hinted at here is real. Currently, there only about a half dozen PWNe with torus plus jet morphologies, with the Crab and (likely) Vela nebulae as the best examples. With a deeper observation, \\pwn\\ may prove to be the third best example. Further work by our group will also include analysis of radio polarization data we have taken with Arecibo. These data may help bolster the geometrical interpretation described here if the swing of the polarization angle across the pulse is measurable. Proper-motion measurements could check if \\psr\\ is moving along the axis of the nebula, which is the case for both Vela and the Crab. However, measuring the proper motion of \\psr\\ (which will likely be small since the distance is likely large) will be very difficult both through radio timing because of timing noise or through VLBI because the radio source is very faint. We are also analyzing VLA observations (in A, C, and D arrays) of the source region to look for a radio PWN and/or SNR. This is important for characterizing the surrounding medium, which has an important effect on the morphology of the PWN and its size." }, "0403/astro-ph0403318_arXiv.txt": { "abstract": " ", "introduction": "The planetary nebula (PN) phase is one of the final stages of stellar evolution for 1-8$\\msun$ stars. A planetary nebula is created when an asymptotic giant branch (AGB) star sheds its hydrogen-rich envelope via stellar wind, exposing the underlying hot carbon-oxygen (CO) core. The surface of the remnant star has a surface temperature T$>$30000K and therefore it emits a significant number of photons with wavelengths, ($\\lambda$), less than 912 Angstroms. These photons with energy greater than 13.6eV are sufficiently energetic to ionize hydrogen in the ejecta. The free electrons collisionally excite the ions in the nebula. The excited atoms then radiatively de-excite creating emission lines. PNe appear green in a small telescope or binoculars because of the [OIII]$\\lambda$5007. The energy input from the star is balanced by the energy emitted by lines from the nebulae. Photons from the star which photoionize the nebulae are the primary energy source. Emission lines carry this energy away. This energy balance determines the temperature of the nebulae and the relative numbers of each ionization stage of each atomic species. PN emission lines can be used to determine to reasonable accuracy the nebular abundances of several atomic species. The density in most PNe is too low for ions to be collisionally de-excited. Therefore, these ions must emit one or more photons to get back to the ground state. By balancing the number of collisional excitations with the number of radiative deexcitations, the abundance of some ionic species can be determined, e.g., using the [OIII]$\\lambda$5007 line we can get the abundance of O$^{+2}$. To get the nebular abundance of an atom, an ionization correction factor is used to account for the unseen ionization stages. Therefore, the uncertainties in each element vary depending on how much of the element is in each stage. The procedure for determining elemental abundances is described in more detail in elementary textbooks on the subject (e.g. Osterbrock 1989). Recently, detailed photoionization models have become widely available, allowing detailed calculations of the nebulae (i.e. CLOUDY). Such models improve our ability to determine nebular abundances. The abundances of several important elements can be inferred for PNe. The list includes hydrogen, helium, carbon, nitrogen, oxygen, neon, sulfur, and argon. Some of these such as neon and argon can not be measured AGB phase which preceeds the PN phase. The gas that makes up a PN has been processed via stellar evolution. This processing leads to the difference between the abundances in PNe and the differences in solar or HII region abundances. The abundances in the Sun and in HII regions are respectively believed to reflect the composition of the interstellar medium (ISM) at 5Gyrs and the current epoch, respectively. The material making up PNe has been processed by stellar evolution. This probably leads to enhancements in the abundances of He, C, and N. The globular cluster M15 contains the well studied planetary nebula (PN) K648. This is one of the few galactic PNe with a reasonably well-determined distance. Therefore, fundamental properties such as the stellar luminosity can be determined with some confidence. Because of its globular cluster membership, many of the progenitor properties, such as the zero age main sequence (ZAMS) mass, can be inferred reliably. Due to the importance of K648 as a halo PN, it has been the focus of several abundance studies, and all of these show the abundances of most metals to be depleted relative to the sun, consistent with a progenitor of low metallicity. Carbon is an exception; studies which determine the ratio (by number) of C/O in K648 infer values that range from $4-11$ (Adams \\etal\\ 1984; Henry, Kwitter, and Howard 1996; Howard, Henry, and McCartney 1997), which is far above C/O in the Sun of 0.43 (\\nocite{ag89}Anders and Grevesse 1989, hereafter AG89). This is in fact much higher than the average C/O ratio of $\\approx 0.8$ for solar neighborhood PNe (\\nocite{rs94}Rola and Stasi{\\'n}ska 1994). Low and intermediate mass stars that have left the main sequence, ascended the giant branch, and passed through the horizontal branch, then enter a thermally unstable phase where energy is generated by shell He and H-burning called the thermally pulsing asymtoptic giant branch (TP-AGB) stage, which is a very important yet not well understood phase [Detailed reviews of this stage can be found in \\nocite{i95} Iben (1995), \\nocite{l93} Lattanzio (1993), and \\nocite{ir83} Iben and Renzini (1983)]. During the TP-AGB stage the star alternates between a long stage where the luminosity is generated mostly by quiescent hydrogen shell burning, with a helium burning layer producing a minority of the energy, and a thermal runaway stage in the unstable helium burning layer (\\nocite{sh65} Schwarzschild and H\\\"{a}rm 1965, 1967; and \\nocite{w66} Weigert 1966). The second stage results in expansion of the outer layers and an extinguishing of the H burning shell. This short stage, characterized by rapid changes, with helium burning dominating the energy generation, is known as a thermal pulse or a He shell flash. TP-AGB stars exhibit large mass-loss rates ranging from 10$^{-7}-$10$^{-4}$ \\msolyr\\ . Indeed such high mass-loss rates are predicted to result in the ejection of the envelope, at which point the star leaves the AGB and becomes a planetary nebula central star (CSPN). The first models of CSPN tracks were made by \\nocite{p71} Paczynski (1971) who showed that the CSPNs evolve horizontally on the HR diagram when nuclear burning is still taking place and then as they cool the luminosity and temperature decrease. \\nocite{hs75} H\\\"arm and Schwarzschild (1975) showed that a CSPN could leave the AGB as either a helium burning or hydrogen burning star. The observational consequences of hydrogen and helium burning have been studied in the more refined models including mass loss showed that the subsequent evolution of the central star depends on whether or not the star leaves the AGB as a helium or hydrogen burner [Sch\\\"{o}nberner (1981, 1983) and \\nocite{i84} Iben (1984)]. Low mass stars (M$\\lesssim$ 3 M$_{\\odot}$) can experience two mixing episodes or ``dredge-ups''. During dredge-up, material that has been processed by nuclear burning is mixed into the surface layers. At the entrance to the giant branch, the convective region can extend into the core, leading to mixing of CNO products into the outer layers. Similarly as shown by \\nocite{i75} Iben (1975), after a thermal pulse on the AGB, the convective region can extend into the core, mixing He-burning products into the outer layers. These two mixing events are known as first and third dredge up, respectively (second dredge up will not concern us here). Therefore, a third dredge-up is a natural explanation of the high carbon abundance found in K648. On the other hand, no carbon stars have been observed either in M15 or in any other globular cluster, although such stars should be the immediate progenitors of objects such as K648 if a third dredge-up occurs. Thus, the lack of carbon stars in M15 weakens the argument for a third dredge-up event. One possible explanation for the absence of carbon stars is a delayed scenario in which the third dredge-up of carbon rich material changes the structure of the envelope during the following interpulse phase, ultimately increasing the mass-loss rate significantly and driving off the stellar envelope (Iben 1995). Thus envelope ejection is delayed until the interpulse phase following this dredge-up of carbon rich material. Another explanation supposes that the envelope is removed during the quiescent He-burning stage that follows a thermal pulse (\\nocite{ren89}Renzini 1989 and \\nocite{rfp88}Renzini and Fuci-Pecci 1988). The carbon then originates in a fast wind from the central star (CSPN). In addition, the wind produces shock-heating in the nebula, which, if not properly accounted for during an abundance analysis, may lead to the inference of a spuriously high C/O ratio. In this case the envelope would be ejected immediately after a thermal pulse while helium shell burning still dominates the luminosity. We refer to this mechanism as the prompt scenario. In this paper we calculate detailed envelope models of thermally pulsing asymptotic giant branch star envelopes to test the predictions of the delayed mechanism, perform other calculations relevant to the prompt mechanism, and compare output of each with observations of K648. Section~2 describes the envelope code, section~3 presents the observational data and the results for the delayed and prompt models, and a brief discussion of our findings is given in section~4. ", "conclusions": "A multivariate data analysis of galactic planetary nebulae has been carried out with the goal of determining the effective dimensionality of PN parameter space as well as the importance of the numerous observational parameters in classifying PNe. The parameters employed for this study are the abundance ratios He/H, N/H, O/H, Ne/H, N/O, Ne/O, and the spatial and kinematic parameters R, Z, and $V_{LSR}$. Our main results are: \\begin{enumerate} \\item Planetary nebula abundance parameter space is two dimensional, i.e. the data can be described by two principal components. The first and most important one is related to the products of stellar nucleosynthesis during the evolution of PN progenitors. The second component is related to progenitor metallicity. \\item The kinematic properties are not linearly correlated with any other parameters. \\item The disk PNe separate into three clusters, each having a distinct location in the nucleosynthesis-metallicity plane. A fourth cluster comprises many of the bulge PNe. The bulge PNe only separate from the disk PNe because of the inclusion of the parameters R and $V_{LSR}$. \\item We have identified some unusual type I PNe with extremely low oxygen abundances. However, it is not clear if these are real. \\item Bulge and disk PNe show the same distributions along the nucleosynthesis and metallicity components. \\end{enumerate} This analysis is a first attempt to characterize planetary nebula parameter space in an objective way. The future addition of parameters such as carbon abundance, nebular expansion velocity, morphology, binarity, and central star mass further improve our understanding of these objects. \\newpage \\chapter{Thermally Pulsing AGB Models} \\label{modelchap} Thermally pulsing asymptotic giant branch (TP-AGB) stars experience two distinct and repeating phases: the helium shell flash or thermal pulse (TP) and the time between pulses or the interpulse phase (IP). The duration of the thermal pulse ($\\tp$) is short compared to the duration of the interpulse phase ($\\tip$) with $\\tp /\\tip\\approx 0.01$. During the TP-AGB, these stars spend most of the time in the interpulse phase, burning both He and H burn quiescently in thin shells. During the interpulse phase, most of the luminosity is produced by the hydrogen burning shell. However, the helium shell in this configuration is unstable, and it will eventually result in a helium shell flash. Thermal pulses occur quasi-periodically, resulting in the luminosity of the helium burning shell increasing by several orders of magnitude. The increased luminosity causes the outer layers of the star to expand, including the hydrogen burning shell. The luminosity generated by hydrogen burning during the shell flash decreases to essentially zero due to the expansion. After the thermal pulse the star resumes its interpulse configuration. This cycle repeats until the envelope is lost via mass-loss. The two distinctive phases of the TP-AGB have important consequences for the surface abundances of the star. At the end of a thermal pulse it is possible for the convective envelope to penetrate into regions where He has been partially burned into $\\ctw$ and minor amounts of $\\oxy$, mixing these products to the stellar surface. This process is called the third dredge-up and is believed to be the process by which carbon stars are produced. In the more massive TP-AGB stars it is possible for the base of the convective envelope to get hot enough to burn $\\ctw$ to $\\cth$ and $\\nit$ and $\\oxy$ to $\\nit$ in a process known as hot-bottom burning. This process is believed to produce the nitrogen rich PNe. Hot bottom burning can take the carbon dredged-up into the envelope and process it into nitrogen. This is one of the possible solutions to the so-called carbon star mystery. The third dredge-up occurs at the end of a thermal pulse and mixes helium and carbon rich material into the stars outer layers. During the thermal pulse a convective shell appears in the helium burning zone. This convection zone spans the mass range from the base of the interpulse He burning shell to just below the position of the convective envelope during the preceding interpulse phase. In this zone these important helium burning reactions take place: \\begin{itemize} \\item{$\\he +\\he\\longrightarrow ^8{\\rm Be}+\\gamma$} \\item{$^8{\\rm Be}+\\he\\longrightarrow\\ctw +\\gamma$} \\item{$\\ctw +\\he\\longrightarrow\\oxy +\\gamma$} \\end{itemize} with the first two being the most important. The third reaction does not occur very often. Therefore, the convective shell at the end of the pulse is composed primarily of helium and carbon. During the transition period between the thermal pulse and the resumption of the interpulse phase the convective shell disappears and the convective envelope can penetrate into this region. The mixing of carbon and helium rich material to the surface layers of the star at the end of a thermal pulse is known as the third dredge-up to distinguish it from the first two dredge-ups which we will describe later." }, "0403/astro-ph0403542_arXiv.txt": { "abstract": "Observations with the SZ effect constitute a powerful new tool for investigating clusters and constraining cosmological parameters. Of particular interest is to investigate how the SZ signal correlates with other cluster properties, such as the mass, temperature and X-ray luminosities. In this presentation we quantify these relations for clusters found in hydrodynamical simulations of large scale structure and investigate their dependence on the effects of radiative cooling and pre-heating. ", "introduction": "} Galaxy clusters are the largest gravitationally bound objects in the Universe. They have typical masses of $10^{14}-10^{15}h^{-1}$~M$_\\odot$ and contain hundreds of galaxies within radius of a few Mpc. The intra-cluster medium (ICM) is filled with hot ionized gas, typically at temperatures 1--15 keV, which produces strong X-ray emission and causes spectral distortions in the CMB spectrum via the Sunyaev--Zel'dovich (SZ) effect (see Ref.~\\cite{sunyaev:1972}). Numerical simulations indicate that the non-baryonic dark matter component in clusters, which is the dominant fraction of their mass, is remarkably self-similar for systems in approximate state of equilibrium, see e.g. Ref.~\\cite{navarro:1997b}. However, the baryonic gas component does not share the same degree of self-similarity. This is more evident from observations of the $L_X$--$T$ relation in clusters, which is much steeper than is predicted by simple self-similar scaling laws, specially for low-mass systems (see e.g. Refs.~\\cite{edge:1991,xue:2000}). This deviation from self-similarity has been interpreted as due to non-gravitational processes, such as radiative cooling and heating, that raise the entropy of the gas, see e.g. Ref.~\\cite{ponman:1999}. The purpose of this study is to investigate the correlation between intrinsic properties of clusters found in simulations, and to compare them with analytical scaling laws. We use high-resolution hydrodynamical simulated clusters to assess the impact of radiative cooling and pre-heating on these scaling relations. Here we focus only on scalings involving the SZ integrated signal, $Y$, at redshift zero. The effects on the mass--temperature and X-ray scaling relations have already been analyzed in detail in Refs.~\\cite{muanwong:2001,muanwong:2002} for this same set of simulations. A more detailed analysis of the SZ scaling relations can be found in Ref.~\\cite{dasilva:2003}, whose results supersede those presented at this conference. ", "conclusions": "We have studied the effect of our models of cooling and pre-heating on the SZ cluster population by identifying clusters in the simulation boxes and computing their characteristic properties. We correlated the SZ luminosity with other cluster properties and derived scaling relations at redshift zero. The non-radiative simulation reproduces well the scalings predicted by the self-similar model, whereas the inclusion of cooling and pre-heating generally changes the slope and normalization of the scaling relations for those runs. The integrated $Y$-signal is found to be tightly correlated with the mass-weighted temperature and total mass of the cluster. The author wishes to thank A. Liddle, P. Thomas, S. Kay and O. Muanwong for many fruitful discussions and acknowledges the use of the computer facilities of the Astronomy Unit at Sussex, UK, and CALMIP at Toulouse, France. The simulations used were carried out on a Cray-T3E at EPCC as part of the VIRGO Consortium collaboration." }, "0403/astro-ph0403068_arXiv.txt": { "abstract": "We present a spectroscopic survey of the giant stellar stream found in the halo of the Andromeda galaxy. Taken with the DEIMOS multi-object spectrograph on the Keck2 telescope, these data display a narrow velocity dispersion of $11\\pm3\\kms$, with a steady radial velocity gradient of $245 \\kms$ over the $125 \\kpc$ radial extent of the stream studied so far. This implies that the Andromeda galaxy possesses a substantial dark matter halo. We fit the orbit of the stream in different galaxy potential models. In a simple model with a composite bulge, disk and halo, where the halo follows a ``universal'' profile that is compressed by the formation of the baryonic components, we find that the kinematics of the stream require a total mass inside $125\\kpc$ of $M_{125} = 7.5^{+2.5}_{-1.3} \\times 10^{11} \\msun$, or $M_{125} > 5.4 \\times 10^{11} \\msun$ at the 99\\% confidence level. This is the first galaxy in which it has been possible to measure the halo mass distribution by such direct dynamical means over such a large distance range. The resulting orbit shows that if M32 or NGC~205 are connected with the stream, they must either trail or lag the densest region of the stream by more than $100\\kpc$. Furthermore, according to the best-fit orbit, the stream passes very close to M31, causing its demise as a coherent structure and producing a fan of stars that will pollute the inner halo, thereby confusing efforts to measure the properties of genuine halo populations. Our data show that several recently identified planetary nebulae, which have been proposed as evidence for the existence of a new companion of M31, are likely members of the Andromeda Stream. ", "introduction": "Stellar streams represent the visible remnants of the merging process by which the halos of galaxies are built up. By studying these streams we can attempt to unravel the formation of galactic halos, seeing when, how and how many small galaxies arrived and were incorporated into large galaxies \\citep{helmi99, johnston01}. Streams are also of great interest as probes of the large-scale mass distribution of the dark halos they reside in \\citep{johnston99, ibata01a, zhao99}. This utility stems from the fact that streams from low-mass disrupting stellar systems trace the orbit of their progenitor, giving a means to constrain the tangential motion of the stars in the stream. The stars must move along the stream, and the magnitude of the tangential velocity must be such that, when the star is integrated along the orbit, it ends up with the same velocity as the stars that are currently downstream on the same orbit. Giant stellar streams may well be common structures around galaxies \\citep{pohlen, malin}. Indeed, the most conspicuous feature in the halo of the Milky Way is the giant rosette stream originating from the Sagittarius dwarf galaxy, which contains approximately half of the high latitude ($|b|>30^\\circ$) intermediate age stars at distances greater than $15\\kpc$ \\citep{ibata01a, ibata02, majewski}. It would appear that the Milky Way has not incorporated into the Halo a more massive galaxy than the Sagittarius dwarf over the last $\\sim 7\\Gyr$. This naturally leads to the question of whether the MW has unusual feeding habits. To answer this, we have undertaken a large photometric study of M31, using the wide field cameras at the INT and CFHT telescopes to resolve stars over the entire disk and inner halo of that galaxy \\citep{ibata01b,ferguson02,mcconnachie}. This has given us an unprecedented panoramic view of the large scale and small scale structure of a disk galaxy. The analysis of this huge dataset is still in progress, but it has already yielded some surprising results regarding the incidence of substructure in the halo of M31. The most prominent of these substructures is a stream-like over-density of stars near the minor axis of M31 \\citep{ibata01b}, at first sight a facsimile of the Sagittarius Stream around the Milky Way. The red giant branch (RGB) stellar density in the halo increases on average by a factor of two in the on-stream regions and is statistically significant at the 50-sigma level. Interestingly, this stream points toward the Andromeda satellites M32 and NGC205, and is aligned with the outer isophotes of NGC205, suggesting a relationship between the Andromeda Stream and these two dwarf galaxies. If this interpretation is correct, the stream has to be the result of previous interactions with M31, as there is otherwise not enough time to spatially separate it from either of the two dwarf galaxies. The proximity of M31 provides us with an opportunity to undertake a spectroscopic survey of individual stars within the stream; indeed Andromeda offers the only extragalactic giant galaxy in which such a study can be undertaken with current instrumentation, as in more distant systems such halo substructure would be smeared into very low surface brightness features. \\begin{figure*} \\ifthenelse{\\UseFigs=1}{ \\includegraphics[angle=270,width=\\hsize]{M31_StreamOrb.fig02.ps}}{} \\caption{For convenience, the left-hand panel reproduces the chart shown previously in Figure~1. The right-hand panel shows a sideways view of the stream, drawn to the same scale as the left-hand panel, which displays the line of sight depth of the fields, together with the positions of M32 and NGC~205. The disk of M31 is highly inclined to our line of sight ($12\\degg5$); the thick line is a schematic representation of a disk of radius $27\\kpc$ inclined at $12\\degg5$. Evidently the stream orbits close to the plane of the Andromeda galaxy.} \\end{figure*} One useful property of the stream is that it is on a highly radial orbit. It passes very close to the centre of M31, where a comparison of the RGB tip of the stream with that of M31 itself shows them to lie at the same distance \\citep{mcconnachie}; whereas in the furthest field that it has been detected to date, the peak of its giant branch is 0.27~mag fainter, indicating that it is $106\\pm20\\kpc$ behind M31 \\citep{mcconnachie}. This fortuitous alignment close to the line of sight, allows us to measure directly the potential gradient over $> 100\\kpc$, and hence measure the halo mass. The layout of this paper is as follows: section~2 presents the spectroscopic survey of the Andromeda Stream, the results of which are used in section~3 to constrain the mass of the dark matter halo of M31, and finally in section~4 we discuss the significance of the results and present the conclusions of this study. Throughout this work, we assume a distance of $780\\kpc$ to M31 \\citep{stanek} and a systemic radial velocity of $-300\\kms$ \\citep{devaucouleurs}. ", "conclusions": "In this first kinematic study of the giant stellar stream in the Andromeda galaxy, we have been able to measure the radial velocity gradient along the stream from the outermost field currently probed, $125\\kpc$ distant from the centre of M31, down to an inner field $\\sim 20\\kpc$ from the centre of that galaxy. Over this huge distance, the (projected) radial velocity changes by $245\\kms$, implying a de-projected velocity difference of $\\sim 280\\kms$. This velocity gradient is used to obtain a zeroth order analytic estimate of the mass of the halo, which for simple halo-only galaxy models such as a logarithmic halo or an NFW halo, implies a mass inside $125\\kpc$ of $M_{125} = 7.6 \\pm 1.2 \\times 10^{11} \\msun$ and $M_{125} = 6.4 \\pm 1.3 \\times 10^{11} \\msun$, respectively. In both cases, the uncertainty associated with a possible tangential velocity of M31 of $300\\kms$, is $\\sim 5$\\%. We also investigate more realistic solutions, allowing the stream to have a non-radial orbit, and taking a galaxy model that is the sum of a disk, bulge and dark halo \\citep{klypin02}. The dark halo of this model is a perturbation on a spherical NFW model, to account for the adiabatic contraction of the dark matter as the baryonic components form. If we disregard the distance data in Fields 12 and 13, since we cannot be certain that the stream is present in those regions, the most likely mass of this galaxy model is $M_{125} = 7.5^{+2.5}_{-1.3} \\times 10^{11}\\msun$, with a lower limit of $M_{125} = 5.4 \\times 10^{11} \\msun$ (at 99\\% confidence). This result is fully consistent with the zeroth order analytic estimates discussed above, and suggests that the derived mass is not very sensitive to the adopted mass model. Furthermore, there is a reasonable agreement of the resulting rotation curve with the kinematics of previously-observed disk tracers (see Figure~6), despite the fact that the model was only fit to the kinematics of the stream. This agreement is due in part to the fact that we have taken previously-fitted models for the disk and bulge, but the halo contribution to the total rotation curve dominates beyond $\\sim 20\\kpc$, and it is at these large distances that we have fitted the model to the stream data. This confers further confidence on the derived mass model. The uncertainty on the mass estimate due to the possible tangential velocity of M31 is likely not very large, approximately 7\\% for a tangential velocity of $300\\kms$. One of the main uncertainties in the present analysis is the flattening of the halo, which we have not explored, as the current data set does not provide sufficient constraints. Future studies, fitting N-body simulations to a larger kinematic sample of stream stars, can be expected to improve the mass estimate and also constrain the halo shape. \\begin{figure} \\ifthenelse{\\UseFigs=1}{ \\includegraphics[angle=0,width=\\hsize]{M31_StreamOrb.fig07.ps}}{} \\caption{The spreading of orbits after a close encounter with M31. The thick-line orbit reproduces the best-fit orbit in the realistic galaxy potential previously shown in Figure~4, integrated from the centre of Field~6. The thin-lines show similar orbits, starting from the same spatial position, but with velocities perturbed by a random offset drawn from a Gaussian distribution of dispersion $11\\kms$. This fanning-out of stars on nearby orbits will lead to the disappearance of the Andromeda Stream, and implies that it is a transient phenomenon. The small square on the minor axis shows the position of the \\citet{brown} ACS field (not drawn to scale!). All other markings are as in Figure~1.} \\end{figure} It is only recently that measurements of the mass of M31 beyond the edge of its gaseous disk have been possible. \\citet{courteau} analysed the velocities of 7 Andromeda satellites, finding a total mass of $13.3\\pm 1.8 \\times 10^{11} \\msun$. In contrast, using a larger sample of 10 satellite galaxies, 17 globular clusters and 9 planetary nebulae as test particles, \\citet{evans00a} found that the most likely total mass of M31 is $12.3^{+18}_{-6} \\times 10^{11}\\msun$, approximately half of their Milky Way estimate of $19^{+36}_{-17} \\times 10^{11}\\msun$. With improved radial velocities of the M31 satellites they were later able to reduce their M31 mass uncertainties, finding a value of the total mass of $\\sim 7.0^{10.5}_{-3.5} \\times 10^{11}\\msun$ \\citep{evans00b}, which is fully consistent with our result of $M_{125} = 7.5^{+2.5}_{-1.3} \\times 10^{11}\\msun$ (defined within a radius of $125\\kpc$). However, a definitive statement of the relative masses of M31 and the Milky Way awaits an improved measurement for the Milky Way. The best fit orbit in the best fit potential is prograde and in the region where it is currently observed, it lies close to the plane of M31. Thus it appears that the orbit of the Andromeda stream is peculiar in being extremely radial, passing very close (within $2\\kpc$) of the centre of M31. This requires very special initial conditions. The stream stars, which are spatially narrowly confined in Fields~1 to 8, will diverge dramatically upon passing close to M31 to form a low-density fan-like structure, since orbits that deviate only slightly from the orbit displayed in Figure~4 on the plunging part of their course will take very different paths after being accelerated around the centre of M31, as demonstrated in Figure~7. The fanning-out of the stream is likely to confuse efforts to measure the metallicity and age of the M31 halo; for instance, the recent discovery by \\citet{brown} of a young halo component in M31 from main-sequence fitting of an extremely deep ACS field may be due to stream contamination (see Figure~7). Thus the stream may be in the process of vanishing as a coherent structure, providing a supply of metal-rich stars into the halo. This also suggests that the stream was removed from its progenitor less than an orbital period ago (the pericentre to pericentre period of the continuous-line orbit in Figure~4 is $1.8\\Gyr$), as we would otherwise not observe the structure as a stream. The ephemeral nature of the stream implies that the progenitor must have survived until $\\sim 1.8\\Gyr$ ago. As we discuss further below, the progenitor was probably of low mass, implying that the rate of decay of its orbit due to dynamical friction was slow, so it followed (or continues to follow) an orbit close to the current orbit of the stream. However, any dwarf galaxy on the derived orbit must have experienced extreme tides as it repeatedly passed close to the centre of M31. One option is that the progenitor was a very dense dwarf galaxy that was sufficiently robust to survive the huge tides. This brings to mind M32 as a candidate, though detailed numerical modeling is required to examine this possibility. The alternative option is that the progenitor of the stream deflected off another halo object, sending it plunging into the current orbit, analogous to the suggestion by \\citet{zhao98} to explain the longevity of the Galactic satellite Sagittarius. However, the connection with M32 presents some difficulties. Although M32 appears to reside in the stream, its velocity is markedly different. For M32 to be associated with the stream would require it to be at a different phase in the orbit (either lagging or trailing). Furthermore, the low velocity dispersion of the stream would appear to preclude M32 as its progenitor (which has $\\sigma_v \\sim 50\\kms$ outside of the nucleus, \\citealt{vandermarel94}), though this cannot be confirmed without a detailed dynamical study. The case for association of NGC~205 with the stream also appears weaker given these kinematic measurements, since the best fit orbit does not overlap with it in phase-space. Future studies may allow us to examine this issue in more detail by following the stream beyond the region currently probed with kinematics. The velocities of the stream stars presented here also shed light on the recent identification of a possible new companion to M31, And VIII \\citep{morrison}. The radial velocities of those planetary nebulae, which are located close to M32 (see Figure~1), appear to have radial velocities consistent with an extrapolation of the Stream, as they lie close to the straight line in the right-hand panel of Figure~1. This would suggest that And VIII is most likely part of the Andromeda Stream, situated in the region of highest over-density reported by \\citet{ibata01b}. However, it is interesting to note that the radial velocities of the \\citet{morrison} planetary nebula sample tend to increase towards the North (showing an apparent positive gradient in the right-hand panel of Figure~1), whereas the Stream stars have a negative velocity gradient. The connection between the two structures therefore merits to be examined more carefully. During the refereeing process, a study by \\citet{merrett} was presented which also investigates the planetary nebulae around M31. However, though their survey has also detected PNe in a region at the base of the stream (near our Field~8), their interpretation is inconsistent with the kinematics of the stars reported here. The direction of motion of the orbit they derive is opposite to ours, and the path of their orbit, which intercepts many M31-disk PNe is substantially different to the orbit that we have fitted. It is possible that their finding reveals the presence of another kinematic structure in Andromeda. The majority of the stream stars that were surveyed have a narrow velocity dispersion of $11\\pm 3\\kms$, though slightly skewed to positive velocities. The fact that our line of sight looks down the stream (see Figure~2), so that we probably see stars over a range of distance along the line of sight, and hence at different phases in the orbit, will tend to render the observed velocity dispersion higher than the intrinsic velocity dispersion. This indicates that the progenitor was most likely a low mass dwarf galaxy. The Milky Way satellite Sagittarius, which has a velocity dispersion of $11\\kms$ \\citep{ibata97}, also has a gigantic stellar stream, but with a larger velocity dispersion of $20\\kms$ \\citep{yanny}. However, it is unclear at present whether the lower velocity dispersion measured in the Andromeda stream compared to the Sagittarius stream implies that its progenitor was of lower mass than Sagittarius, or not." }, "0403/astro-ph0403297_arXiv.txt": { "abstract": "{We present integrated colours, integrated spectral energy distributions, and absorption-line indices, for instantaneous burst solar-metallicity binary stellar populations with ages in the range $1 - 15\\,$Gyr. By comparing the results for populations with and without binary interactions we show that the inclusion of binary interactions makes the appearance of the population substantially bluer -- this is the case for each of the quantities we have considered. This effect raises the derived age and metallicity of the population. Therefore it is necessary to consider binary interactions in order to draw accurate conclusions from evolutionary population synthesis work. ", "introduction": "The majority of current studies of the integrated light of stellar populations using the evolutionary population synthesis (EPS) method -- the most direct approach for modelling galaxies -- have tended to focus solely on the evolution of single stars. However, observations tell us that upwards of 50\\% of the stars populating galaxies are expected to be in binary or higher-order multiple systems (Duquennoy \\& Mayor 1991; Richichi et al. 1994, for example). Binary evolution, if the component stars are close enough to exchange mass, can drastically alter the evolution path of a star as expected from single star evolution. Binary interactions can also create some important classes of objects, such as blue stragglers (BSs: Pols \\& Marinus 1994), and subdwarf B stars (sdBs, also referred as extreme horizontal branch [EHB] stars: Han et al. 2002; 2003). Therefore binary stars have the potential to play an important role in determining the overall appearance of any realistic stellar population and their evolution should be accounted for in population synthesis models. A few EPS studies have previously made advances in this direction (e.g. Pols \\& Marinus 1994; Cervino et al. 1997) but only for specialised cases and not to the degree that we present here. In this letter we assume that all stars are born in binaries and born at the same time, i.e. an instantaneous binary stellar population (BSP). We then model any interactions within these binaries in our EPS models to investigate the effects on the integrated colours, integrated spectral energy distributions (ISEDs), and absorption feature indices. The outline of the paper is as follows: we describe our EPS models and algorithm in Sect. 2, present our results and some discussion in Sect. 3, and then finally in Sect. 4 we give our conclusions. ", "conclusions": "We have simulated realistic stellar populations composed of 100\\% binaries by producing $2 \\times 10^5$ binary systems using a Monte Carlo technique. We computed the integrated colours, ISEDs and Lick/IDS absorption feature indices for these instantaneous burst BSPs with and without binary interactions. In comparison we find that modelling binary evolution, and the additional classes of stars that this produces, leads to bluer integrated colours, ISEDs and Lick/IDS absorption feature indices, and therefore makes the derived age and metallicity raise. In this letter we have only considered the effects of binaries for solar metallicity BSPs -- more detailed studies will be given later." }, "0403/astro-ph0403012_arXiv.txt": { "abstract": "New precision in measuring extragalactic distances using supernovae has confirmed with high probability an accelerating increase in redshift with distance. This has been interpreted as implying the existence of dark energy in an expanding and accelerating, flat universe. A more logical explanation of these observations follows directly from an observation made by Erwin Schr\\\"odinger in 1939 that in a closed Friedmann universe $every$ quantum wave function changes with spacetime geometry. Double the size of the universe and both the wavelengths of photons and the sizes of atoms double. When the evolution of atoms and photons are combined, the meaning of Hubble redshift is reversed. Redshift is characteristic of contracting universes. The magnitude-redshift curve for a contracting universe has exactly the accelerating form recently observed and is in excellent quantitative agreement with the data of Riess et al. 1998, Knop et al. 2003, and others. An observed maximum redshift of 1.3 gives a minimum age estimate for the universe of $114$ billion years. The time until collapse is estimated to be 15 billion years or less. ", "introduction": "Supernovae provide an excellent opportunity for measuring extragalactic distances. The observations of these exploding stars using the best earth and space-based telescopes, a physical understanding of supernovae evolution, and sophisticated analyses of their spectra have led to the surprising conclusion that the expansion of the universe seems to be accelerating (Perlmutter \\& Schmidt 2003). The upward trend of the magnitude versus redshift curve at high redshift is clear in Figure 1 (Riess et al. 1998). This departure from a linear or decelerating trend has been explained as eternal expansion with a positive cosmological constant and a current acceleration of expansion. This conclusion is based on a conventional interpretation of redshift as the result of photon wavelength increasing in an expanding Friedmann universe. This was one idea originally suggested by Hubble \\& Tolman (1935), but it is interesting to note that they concluded their paper with \\begin{quote} It also seemed desirable to express an open-minded position as to the true cause of the nebular red-shift, and to point out the indications that spatial curvature may have to play a part in the explanation of existing nebular data. \\end{quote} The conclusion drawn here is that spatial curvature indeed plays a critical part in explaining cosmic redshift. It is the combination of changes in photons during their transit time and changes in atoms between the time of photon emission and later when they provide the wavelength standard, that defines redshift. ", "conclusions": "Recent supernovae data may be interpreted as providing experimental confirmation of the theoretical predictions of the coevolution of spacetime geometry and quantum wave functions made by Schr\\\"odinger (1939), Sumner (1994), and Sumner \\& Sumner (2000). This explanation of the magnitude-redshift relationship comes directly and unambiguously from the roots of the best-tested physical theories of the 1930's. \\bigskip \\medskip The author thanks Dawn Sumner for many insights and helpful suggestions. \\newpage" }, "0403/astro-ph0403224_arXiv.txt": { "abstract": "We construct models of core-collapse supernovae in one spatial dimension, including rotation, angular momentum transport, and viscous dissipation employing an $\\alpha$-prescription. We compare the evolution of a fiducial 11\\,M$_\\odot$ non-rotating progenitor with its evolution including a wide range of imposed initial rotation profiles ($1.25