{ "0808/0808.3221_arXiv.txt": { "abstract": "The precise localization of short/hard (Type I) gamma-ray bursts (GRBs) in recent years has answered many questions but raised even more. I present some results of a systematic study of the optical afterglows of long/soft (Type II) and short/hard (Type I) GRBs, focusing on the optical luminosity as another puzzle piece in the classification of GRBs. ", "introduction": "In the last decade, since the precise localization via the discovery of their optical \\citep{vanParadijs970228}, X-ray \\citep{Costa970228} and radio \\citep{Frail970508} afterglows, much knowledge has been gained about the high-energy transients known as gamma-ray bursts, which we now know are extremely energetic \\citep{Kulkarni990123} explosions of massive stars \\citep{WoosleyBloom} in cosmological distances \\citep{Metzger970508}. But there are two classes of GRBs \\citep{Kouveliotou1993}, and those of short duration remained enigmatic until the advent of the dedicated GRB mission \\emph{Swift} \\citep{GehrelsSwift}. This satellite made it possible to also pinpoint short GRBs via their afterglows in the X-ray \\citep{Gehrels050509B}, optical \\citep{Hjorth050709, Fox050709} and radio \\citep{Berger050724} regimes, establishing not only that they too originate at cosmological distances, but also that at least some of them are not associated with recent star formation \\citep{Berger050724, Barthelmy050724}, implying that they must originate from different progenitors than long/soft GRBs. The most favored model is the coalescence of two compact objects of which at least one is a neutron star \\citep[for reviews, see][]{NakarReview, LeeRR}. The discovery of two temporally long GRBs with no associated supernova emission down to deep limits \\citep{Fynbo060614, Gal-Yam060614, DellaValle060614, Ofek060505} exacerbated the debate on GRB classification \\citep{Gehrels060614} which had been triggered by GRB 050724, a GRB in an elliptical galaxy which was temporally long as seen by \\emph{Swift}, but would have been classified as short if detected by BATSE, due to the presence of a long, soft, X-ray-bright tail of emission. This led \\cite{Zhang060614}, in the light of classifying GRBs according to the underlying progenitor physics \\citep{BloomPhysics}, to create a new classification based on several observed quantities and phenomena, such as duration, spectral lag \\citep{NorrisBonnell}, light curve shape and host galaxy characteristics. GRBs that originate in the explosions of massive stars are called Type II events, whereas those unassociated with recent star formation (probably due to compact object coalescence) are Type I GRBs. By the end of 2007, the \\emph{Swift} mission had resulted in a plethora of (optical) afterglow discoveries, allowing the creation of a large sample of Type II GRB afterglows, as well as a first moderately large sample of Type I GRB afterglows. Since several years, I and my collaborators have been systematically studying large samples of optical afterglows. In this article, I report results from two recent papers, the first one \\citep{KannPaperIV} comparing the afterglows of \\emph{Swift}-era Type II GRBs with those of the pre-\\emph{Swift} era \\citep{KannPaperIII}, the second one comparing the afterglows of Type I GRBs to the complete Type II GRB afterglow sample \\citep{KannPaperV}. I focus especially on three enigmatic events, GRBs 060121, 060505 and 060614, which point to problems in the Type I/II classification scheme, in the light of the luminosities of their optical afterglows. ", "conclusions": "I have presented the results of two studies on Type I and Type II GRB afterglows in the \\emph{Swift} era, with a special focus on three enigmatic and hard-to-classify GRBs in the light of their optical afterglow luminosities. In none of the three cases is the afterglow luminosity a decisive factor in determining the GRB progenitor class, but they still offer a further puzzle piece toward a deeper understanding of GRBs and their progenitors. \\begin{theacknowledgments} D.A.K. acknowledges B. Zhang for the invitation to the Nanjing conference as well as comments, S. Klose for all the financial support, and the contribution of the many authors who had a part in creating the two papers that this proceeding is based upon. Also thanks to Y. F. Huang for many an excellent information both before and after the conference. \\end{theacknowledgments} \\doingARLO[" }, "0808/0808.3401_arXiv.txt": { "abstract": "The abundance and structure of dark matter subhalos has been analyzed extensively in recent studies of dark matter-only simulations, but comparatively little is known about the impact of baryonic physics on halo substructures. We here extend the {\\small SUBFIND} algorithm for substructure identification such that it can be reliably applied to dissipative hydrodynamical simulations that include star formation. This allows, in particular, the identification of galaxies as substructures in simulations of clusters of galaxies, and a determination of their content of gravitationally bound stars, dark matter, and hot and cold gas. Using a large set of cosmological cluster simulations, we present a detailed analysis of halo substructures in hydrodynamical simulations of galaxy clusters, focusing in particular on the influence both of radiative and non-radiative gas physics, and of non-standard physics such as thermal conduction and feedback by galactic outflows. We also examine the impact of numerical nuisance parameters such as artificial viscosity parameterizations. We find that diffuse hot gas is efficiently stripped from subhalos when they enter the highly pressurized cluster atmosphere. This has the effect of decreasing the subhalo mass function relative to a corresponding dark matter-only simulation. These effects are mitigated in radiative runs, where baryons condense in the central subhalo regions and form compact stellar cores. However, in all cases, only a very small fraction, of the order of one percent, of subhalos within the cluster virial radii preserve a gravitationally bound hot gaseous atmosphere. The fraction of mass contributed by gas in subhalos is found to increase with the cluster-centric distance. Interestingly, this trend extends well beyond the virial radii, thus showing that galaxies feel the environment of the pressurized cluster gas over fairly large distances. The compact stellar cores (i.e.~galaxies) are generally more resistant against tidal disruption than pure dark matter subhalos. Still, the fraction of star-dominated substructures within our simulated clusters is only $\\sim 10$ per cent. We expect that the finite resolution in our simulations makes the galaxies overly susceptible to tidal disruption, hence the above fraction of star--dominated galaxies should represent a lower limit for the actual fraction of galaxies surviving the disruption of their host dark matter subhalo. ", "introduction": "\\label{sec:intro} The hierarchical cold dark matter (CDM) model has been established as the standard model of cosmic structure formation and its parameters are now tightly constrained by a variety of observations \\citep[e.g.][\\ \\ and references therein]{2008arXiv0803.0547K, 2006Natur.440.1137S}. Within the CDM scenario, galaxy clusters are the largest gravitationally bound objects in the universe, and they exhibit rich information about the process of structure formation. As a result, they attract great interest both from observational and theoretical points of view. Due to their recent formation, clusters of galaxies also represent the ideal places where the hierarchical assembly of structures is `caught in the act'. Indeed, the complexity of their internal structure reflects the infall of hundreds, if not thousands, of smaller objects and their subsequent destruction or survival within the cluster potential. The dynamical processes determining the fate of the accreting structures also provide the link between the internal dynamics of clusters and the properties of their galaxy population. Modern cosmological simulations provide an ideal tool to study the non-linear dynamics of these processes in a cosmological environment. Thanks to the high resolution reached by dark matter (DM) only simulations, a consistent picture of the population of subhalos within galaxy clusters has now emerged \\citep[e.g., ][]{1999ApJ...524L..19M,2000ApJ...544..616G,2001MNRAS.328..726S,2002MNRAS.335L..84S, 2003MNRAS.345.1313S,2004MNRAS.352..535D,2004MNRAS.352L...1G,2004MNRAS.348..333D,2004ApJ...609...35K}. Although these simulations give a highly detailed description of the evolution of DM substructures, they can not provide information on how gas-dynamical processes affect the properties and the dynamics of substructures. Yet such an understanding is needed to make a more direct link of the properties of dark matter substructures with those of the galaxies that orbit in clusters. For example, we would like to know the timescale over which subhalos can retain their diffuse gas once they infall into a cluster, as this also determines the time interval over which gas can continue to cool and to refuel ongoing star formation. Being highly pressurized, the intra--cluster medium (ICM) is quite efficient in stripping gas from merging structures (\\citealt{1972ApJ...176....1G}; see also \\citealt{2008MNRAS.383..593M} and references therein), thereby altering both the mass of the subhalos and their survival time. The stripping also governs the timescale over which galaxy morphology and colours are modified inside galaxy clusters \\citep{1999MNRAS.308..947A,2004AJ....127.3361K}. Finally, although baryons are expected to only play a sub-dominant role for the dynamics of substructures due to their limited mass fraction, they may nevertheless alter the structure of subhalos in important ways, affecting their survival, mass loss rate, abundance and radial distribution. Quantifying these differences relative to dark matter-only simulations is an important challenge for the present generation of cosmological hydrodynamical simulations. So far there have only been a limited number of investigations of the properties of sub-halos within hydrodynamical simulations. Based on cluster simulationsw with the adaptive mesh refinement code ART, \\citet{2005ApJ...618..557N} have demonstrated that the stellar mass of infalling galaxies is approximately conserved and the survival of the galaxies is slightly enhanced. Similar studies using the Tree-SPH code GASOLINE have been performed by \\citet{2006MNRAS.366.1529M}, where twice as many sub-halos within the central part of a Galaxy sized halo were found when compared with pure dark matter simulations. Finally \\citet{2008ApJ...678....6W} used a TreeSPH code to investigate the halo occupation and sub-halo correlation functions for one group-sized halo, also pointing out the enhanced survival of sub-structures in the hydrodynamical runs. Here we present a systematic analysis of a large sample of galaxy clusters, with an emphasis on the role of non-standard physics and the details of feedback prescriptions on the survival of galaxies. Additionally, we also present an analysis of the evolution of the diffuse gas within the galaxies. In general our findings are in broad agreement with those in previous studies, but we stress that our star dominant systems are still not as numerous as expected from semi-analytic modelling to reproduce the cluster luminosity functions (e.g., \\citet{2007MNRAS.375....2D}; see also \\citet{2006RPPh...69.3101B} for a review on semi-analytic models of galaxy formation). Thanks to improvements in numerical codes and higher computer performance, cosmological hydrodynamical simulations of galaxy clusters can now model a number of the most important physical processes in galaxy formation, such as radiative cooling, star formation, and feedback in energy and metals, while at the same time reaching sufficient resolution to treat these processes in a numerically robust and physically meaningful way \\citep[e.g., ][\\ , for a recent review]{2008SSRv..134..269B}. This allows more realistic structure formation calculations of the coupled system of dark matter and baryonic gas than possible with dark matter only simulations, far into the highly non-linear regime. For instance, radiative cooling has the effect of letting a significant fraction of the baryons cool at the halo centres, which alters the shape and concentration of halos through the mechanism of adiabatic contraction \\citep[e.g.,][]{2004ApJ...616...16G}. At the same time, the condensation of cooled baryons and their subsequent conversion into collisionless stars is expected to largely protect them from stripping, reducing the rate of disruption of substructures and increasing their survival times. Indeed, assuming that galaxies at least survive for a while the disruption of their host DM halos has been shown to be crucial for semi--analytical models of galaxy formation to provide a successful description of the observed galaxy population in clusters \\citep[e.g., ][]{2004MNRAS.352L...1G}. This discussion highlights the importance of understanding the properties and the evolution of cluster substructures as predicted by modern cosmological hydrodynamical simulations. Evidently, an important technical prerequisite for such an analysis is the availability of suitable numerical algorithms to reliably identify substructure in dissipative simulations. To this end we introduce a modified version of the {\\small SUBFIND} algorithm, and test its robustness. We then apply it in a detailed analysis of the properties of substructures in a large ensemble of galaxy clusters, which have been simulated both at different resolutions in their DM-only version, and in several further versions where different physical processes where included, such as non-radiative and radiative hydrodynamics, star formation, energy feedback, gas viscosity and thermal conduction. The aim of this analysis is to quantify the numerical robustness of the measured properties of substructures, and to identify the quantitative influence of these physical processes on substructure statistics in comparison with dark matter only models. The paper is organized as follows. We describe in Section~\\ref{sec:sim} the sample of galaxy clusters and the simulations performed. Section~\\ref{sec:detection} provides a description of the algorithm used to identify gravitationally bound substructures. This algorithm is a modified version of SUBFIND \\citep{2001MNRAS.328..726S}, which we suitably changed to allow extraction of bound structures from N-Body/SPH simulations also in the presence of gas and star particles. In Section~\\ref{sec:sub}, we present our results about the properties of the substructures in both DM and hydrodynamical simulations. We summarize our results and outline our main conclusion in Section \\ref{sec:conc}. An Appendix is devoted to the presentation of tests of resolution and numerical stability of our analysis. ", "conclusions": "\\label{sec:conc} In this study, we presented an analysis of the basic substructure properties in high--resolution hydrodynamical simulations of galaxy clusters carried out with the TreePM-SPH code {\\small GADGET}. We used a fairly large set of 25 clusters, with virial masses in the range $1-33 \\times 10^{14}\\msun$, out of which eight clusters have masses above $10^{15}\\msun$. The identification of the substructures has been carried out with the {\\small SUBFIND} algorithm \\citep{2001MNRAS.328..726S}, which we suitably modified to account for the presence of gas and star particles in hydrodynamical simulations. The primary aim of our analysis was to quantify the impact that gas physics has on the properties of subhalos, especially on their abundance. Note that previous work has so far almost exclusively analyzed substructures in dark matter-only simulations, so that our study is one of the first attempts to directly compare the DM-only results to those of hydrodynamical simulations that also account for baryonic physics. We analyzed non--radiative hydrodynamical simulations as well as radiative simulations, including star formation and different efficiencies for energy feedback associated with galactic outflows. Also, we examined the sensitivity of our results with respect to different parameterizations for the artificial viscosity needed in SPH. The main results of our analysis can be summarized as follows. \\begin{itemize} \\item Consistent with previous studies \\citep{2001MNRAS.328..726S,2004MNRAS.348..333D}, we find that the subhalo mass function in DM-only runs converges well for different numerical resolution over the mass range where substructures are resolved with at least $\\simeq 30$ gravitationally bound DM particles. We extend this result to show that the same convergence also holds well for the total subhalo mass function in hydrodynamical runs, irrespective of whether cooling and star formation are included in the simulations. \\item In non--radiative hydrodynamical simulations, the presence of a gas component leads to a reduction of the total subhalo mass function with respect to DM-only runs. On average, the reduction in mass of the substructures is slightly larger than expected from the complete stripping of the baryonic component. This is due to a combination of the modification of the subhalo orbits arising from the pressure force exerted by the intra--cluster medium. Furthermore, we see indications for an increased susceptibility of the surviving DM halos to tidal disruption. \\item In general we find that only a very small fraction of subhalos, of order of one percent, within the cluster virial radii maintain a self-bound atmosphere of hot gas. This indicates a high efficiency for the stripping of the hot gas component in our simulations. The radial dependence of the mean gas fraction within subhalos indicates that gas stripping starts already at large cluster--centric distances, beyond the virial radius. Using a scheme for reduced artificial viscosity has the effect of reducing viscous stripping in the cluster periphery, while increasing the stripping within the virialized high-density region. \\item In radiative runs that include star formation, some of the baryons are protected from stripping due to their concentration at the centres of the subhalos. In this case, substructures have masses which are comparable to, or are slightly larger than in the DM-only runs. The stellar mass fraction is also found to strongly increase for subhalos close to the centre. This confirms that compact clumps of star particles are indeed more resistant against tidal destruction than DM subhalos. Despite this, we find that only a small fraction of subhalos are dominated by their stellar component. As a result, the number of star--dominated galaxies is smaller than expected when, like usually assumed in semi--analytical models of galaxy formation, they are allowed to survive for a dynamical friction time after the disruption of their DM halo. This suggests that our hydrodynamical simulations at their current resolution may not be able yet to accurately follow the survival of galaxies within clusters after the destruction of their DM subhalos. \\end{itemize} Our study shows that the technical improvements we implemented in {\\small SUBFIND} allow a reliable identification of substructures in hydrodynamical simulations that include radiative cooling and star formation. This allows a direct identification of satellite galaxies not only in clusters of galaxies, but also in future high-resolution hydrodynamical simulations of galaxy-sized halos. This is essential to make direct contact with observational data, and also allows studies of the structural impact of baryonic physics on substructure. As our result show here, gas physics {\\em does alter} substructure statistics in interesting ways, but the effects depend on physical processes such as radiative cooling and feedback. This also means that the approximate self-similarity of substructure properties that has been found in DM-only simulations of clusters and galaxy-size objects \\citep{1999ApJ...524L..19M} is not expected to hold nearly as well in simulations with radiative cooling. For example, in galaxy-size halos, many subhalos will be so small that they do not experience atomic cooling; as a result their behaviour should be close to our non-radiative results, and we would here expect a reduction of their abundance relative to the DM-only case. In future work, it will be important to further improve the numerical description and resolution of our simulations, in order to be able to more accurately follow in particular the star--dominated substructures (i.e., galaxies) within galaxy clusters. This will then also provide important input for the semi--analytical models of galaxy formation, allowing a check and calibration of their dynamical friction and gas stripping prescriptions." }, "0808/0808.3771_arXiv.txt": { "abstract": "\\noindent We present analyses of a 50\\,ks observation of the supergiant X-ray binary system \\ifApJ{Cygnus\\,X-1 (\\mbox{Cyg\\,X-1})/\\linebreak}{Cygnus\\,X-1/}HDE\\,226868 taken~with the \\Chandra{} High Energy Transmission Grating Spectrometer (HETGS). Cyg\\,X-1 was in its spectrally hard state and the observation was performed during superior conjunction of the black hole, allowing for the spectroscopic analysis of the accreted stellar wind along the line of sight. A significant part of the observation covers X-ray dips as commonly observed for \\Cyg{} at this orbital phase, however, here we analyze only the high count rate nondip spectrum. The full 0.5--10\\,keV continuum can be described by a single model consisting of a disk, a narrow and a relativistically broadened Fe K$\\alpha$ line, and a power-law component, which is consistent with simultaneous \\ifApJ{\\textsl{Rossi X-Ray Timing Explorer}}{\\RXTE} broad band data. We detect absorption edges from overabundant neutral O, Ne, and Fe, and absorption line series from highly ionized ions and infer column densities and Doppler shifts. With emission lines of He-like Mg\\,\\textsc{xi}, we detect two plasma components with velocities and densities consistent with the base of the spherical wind and a focused wind. A simple simulation of the photoionization zone suggests that large parts of the spherical wind outside of the focused stream are completely ionized, which is consistent with the low velocities ($<$200\\,km\\,s$^{-1}$) observed in the absorption lines, as the position of absorbers in a spherical wind at low projected velocity is well constrained. Our observations provide input for models that couple the wind activity of HDE\\,226868 to the properties of the accretion flow onto the black hole. ", "introduction": "\\label{sec:intro} \\object[Cygnus X-1]{Cygnus\\,X-1\\ifApJ{ (Cyg X-1)}{}} was discovered in 1964 \\citep{Bowyer1965} and soon identified as a high-mass X-ray binary system (HMXB) with an orbital period of 5.6\\,d \\citep{MurdinWebster1971,WebsterMurdin1972,Bolton1972}. It consists of the supergiant O9.7 star \\object[HDE 226868]{HDE\\,226868} \\citep{Walborn1973,Humphreys1978} and a compact object, which is dynamically constrained to be a black hole \\citep{GiesBolton1982}. The detailed spectroscopic analysis of HDE\\,226868 by \\citet{Herrero1995} gives a stellar mass $M_\\star\\approx18\\,M_\\odot$, leading to a mass of $M_\\mathrm{BH}\\sim$10\\,$M_\\odot$ for the black hole, if an inclination $i\\approx35^\\circ$ is assumed. Note that \\citet{Ziolkowski2005} derives a mass of $M_\\star=(40\\pm5)\\,M_\\odot$ from the evolutionary state of HDE\\,226868, corresponding to $M_\\mathrm{BH}=(20\\pm5)\\,M_\\odot$, while \\citet{Shaposhnikov2007} claim $M_\\mathrm{BH}=(8.7\\pm0.8)\\,M_\\odot$ from X-ray spectral-timing relations. \\Cyg{} is usually found in one of the two states that are distinguished by the soft X-ray luminosity and spectral shape, the timing properties, and the radio flux \\citep[see, e.g.,][]{Pottschmidt2003,Gleissner2004_III,Gleissner2004_II,Wilms2006}: the low/hard state is characterized by a lower luminosity below 10\\,keV, a hard Comptonization power-law spectrum (photon index $\\Gamma\\sim1.7$) with a cutoff at high energies (folding energy $E_\\mathrm{fold}\\sim150\\,$keV) and strong variability of $\\sim$30\\% root mean square (rms). Radio emission is detected at the $\\sim$15\\,mJy level. In the high/soft state, the soft X-ray spectrum is dominated by a bright and much less variable (only few~\\% rms) thermal disk component, and the source is invisible in the radio. Within the classification of \\citet{RemillardMcClintock2006}, the high/soft state of \\Cyg{} corresponds to the steep power-law state rather than to the thermal state, as a power-law spectrum with photon index $\\Gamma\\sim2.5$ may extend up to $\\sim$10\\,MeV \\citep{Zhang1997,McConnell2002,CadolleBel2006}. Most of the time, \\Cyg{} is found in the hard state, but transitions to the soft state and back after a few weeks or months are common every few years. Transitional or intermediate states \\citep{Belloni1996} are often accompanied by radio and/or X-ray flares. Similar to a transition to the soft state, the spectrum softens during these flares and the variability is reduced. This behavior is called a ``failed state transition'' if the true soft state is not reached \\citep{Pottschmidt2000,Pottschmidt2003}. Transitional states have occurred more frequently since mid-1999 than before \\citep{Wilms2006}, which might indicate changes in the mass-accretion rate due to a slight expansion of HDE\\,226868 \\citep{Karitskaya2006}. HMXBs are believed to be powered by accretion from the stellar wind. The accretion rate and therefore X-ray luminosity and spectral state are thus very sensitive to the wind's detailed properties such as velocity, density, and ionization. For HDE\\,226868, \\citet{Gies2003} found an anticorrelation between the H$\\alpha$ equivalent width (an indicator for the wind mass loss rate $\\dot M_\\star$) and the X-ray flux. Considering the photoionization of the wind would allow for a self-consistent explanation \\citep[see, e.g.,][]{Blondin1994}: a \\emph{lower} mass loss gives a \\emph{lower} wind density and therefore \\emph{higher} degree of ionization due to the irradiation of hard X-rays, which reduces the driving force of HDE\\,226868's UV photons on the wind and results in a \\emph{lower} wind velocity $v$, leading finally to a \\emph{higher} accretion rate \\citep[$\\propto\\dot M_\\star/v^4$,][]{BondiHoyle1944}. However, \\citet{Gies2008} find suggestions that the photoionization and velocity of the wind might be similar during both hard and soft states. UV observations allow the photoionization in the HDE\\,226868\\,/\\,\\Cyg{} system to be probed: \\citet{Vrtilek2008} reported P\\,Cygni profiles of \\Ion{N}{v}, \\Ion{C}{iv}, and \\Ion{Si}{iv} with weaker absorption components at orbital phase $\\phi_\\mathrm{orb}\\approx0.5$, i.e., when the black hole is in the foreground of the supergiant. This reduced absorption, which was already found by \\citet{Treves1980}, is due to the \\citet{HatchettMcCray1977} effect, showing that those ions become superionized by the X-ray source. \\citet{Gies2008} model the orbital variations of the UV lines assuming that the wind of HDE\\,226868 is restricted to the shadow wind from the shielded side of the stellar surface \\citep{Blondin1994}, i.e., the \\citet{Stromgren1939} zone of \\Cyg{} extends to the donor star. However, this assumption applies only to the spherical part of the wind, which might therefore hardly contribute to the mass accretion of \\Cyg{}. As HDE\\,226868 is close to filling its Roche lobe \\citep{Conti1978,GiesBolton1986_II,GiesBolton1986_III}, the wind is not spherically symmetrical as for isolated stars, but strongly enhanced toward the black hole \\citep[``focused wind\\ifApJ{;''}{'';}][]{FriendCastor1982}. The strongest wind absorption lines in the optical are therefore observed at the conjunction phases \\citep{Gies2003}. Similarly, X-ray absorption dips occur preferentially around $\\phi_\\mathrm{orb}=0$, i.e., during superior conjunction of the black hole \\citep{BalucinskaChurch2000}. These dips are probably caused by dense, neutral clumps, formed in the focused wind where the photoionization is reduced, although recent analyses have also suggested that part of the dipping activity may result from the interaction of the focused wind with the edge of the accretion disk \\citep{Poutanen2008}. The photoionization and dynamics of both the spherical and focused winds can also be investigated with the high-resolution grating spectrometers of the modern X-ray observatories \\Chandra{} or \\textsl{XMM-Newton}. As none of the previously reported observations of \\Cyg{} was performed at orbital phase $\\phi_\\mathrm{orb}=0$ and in the hard state, when the wind is probably denser and less ionized than in the soft state, the \\Chandra{} observation presented here allows for the most detailed investigation of the focused wind to date. The remainder of this paper is organized as follows: in \\Sect{sec:obs}, we describe our observations of \\Cyg{} with \\Chandra{} and the \\textsl{Rossi X-Ray Timing Explorer} (\\RXTE), and how we model CCD pile-up for the \\Chandra-% \\ifApJ{High Energy Transmission Grating Spectrometer (HETGS)}{HETGS} data. We present our investigations in \\Sect{sec:analysis}: after investigating the light curves, we model the nondip continuum and analyze neutral absorption edges and absorption lines from the highly ionized stellar wind -- and the few emission lines from He-like ions, which indicate two plasma components. In \\Sect{sec:discuss}, we discuss models for the stellar wind and the photoionization zone. We summarize our results after comparing them with those of the previous \\Chandra{} observations of \\Cyg. ", "conclusions": "\\label{sec:discuss} In the following, we discuss our results and derive constraints on the stellar wind in the accretion region. \\subsection{Velocity and Density of the Stellar Wind}\\label{sec:winddensity} While a spherically symmetric model for the stellar wind in the HDE\\,226868/\\Cyg{} system can be excluded by observations \\cite[see, e.g.,][]{GiesBolton1986_III,Gies2003,Miller2005,Gies2008}, a symmetric velocity law \\begin{equation} v(r) \\;\\;=\\;\\; v_\\infty \\;\\cdot\\; f(r/R_\\star) \\label{eq:vr} \\end{equation} is usually assumed to obtain a first estimate of the particle density in the wind. The fraction $f$ of the terminal velocity $v_\\infty$ is often parameterized by \\citep[\\Equation~3]{LamersLeitherer1993} \\begin{equation} f(x) \\;\\;=\\;\\; f_0 \\:\\;+\\:\\; \\left(1-f_0\\right) \\cdot \\left(1-1/x\\right)^\\beta \\quad (\\mbox{for } x\\ge1) \\label{eq:f} \\end{equation} with $f_0:=v_0/v_\\infty$, where $v_0$ is the velocity at the base of the wind ($x=1$). The simple model for the radiatively driven wind of isolated stars by \\citet{CastorAbbottKlein1975} is obtained for $\\beta=1/2$. The photoionization of the wind, however, suppresses its acceleration \\citep{Blondin1994}, such that a smaller $f$ (e.g., a larger $\\beta$ within the same model) is required in the Str\\\"omgren zone. \\citet{GiesBolton1986_III} have explained the orbital variation of the 4686\\,\\AA{} He\\,\\textsc{ii} emission line profile of \\Cyg{} with a similar model for the focused wind, where $v_\\infty$, $R_\\star$, and $\\beta$ depend on the angle~$\\theta$ from the binary axis. $\\beta$~was interpolated between 1.60 and 1.05 for $\\theta$ between $0$ and $20^\\circ$. The value $\\beta(\\theta\\!=\\!20^\\circ)=1.05$ is, however, often used for a spherically symmetrical wind as well \\citep{Lachowicz2006,Szostek2007,Poutanen2008}. \\citet{Vrtilek2008} use $f_0=0.01$ to avoid numerical singularities and fit the (relatively low) value $\\beta\\approx0.75$ to their models for UV lines. $v_0$ is likely to be of the order of the thermal velocity of H atoms, which is $(2kT/m_\\mathrm{H})^{1/2}=23\\,$km\\,s$^{-1}$ and corresponds to $f_0=0.011$ for HDE\\,226868's effective temperature $T_\\mathrm{eff}=32$\\ifApJ{,}{\\,}000\\,K and $v_\\infty=2100\\,$km\\,s$^{-1}$ \\citep{Herrero1995}. \\begin{table}\\centering \\caption{Solutions to \\Eq{eq:nr_numbers} for the Model of \\Eq{eq:f}} \\label{tab:density_solutions} \\begin{tabular}{cccccc} \\hline \\hline $n_\\mathrm{H}(x)$ & $d(x)$ & $f_0$ & $\\beta$ & $x$ & $2100 \\cdot f(x)$ \\\\ (10$^{10}$\\,cm$^{-3}$) & & $=v_0/v_\\infty$ & & $=r/R_\\star$ & for $\\beta=\\beta_\\mathrm{max}$ \\\\ \\hline \\eVS 440 & 200 & 0.005 & $<\\infty$ & 1 & 11\\\\ \\hline \\eVS 110 & 50 & 0.011 & $\\le1$ & $\\le1.01$ & $\\le41$\\\\ 110 & 50 & 0.011 & $\\le2$ & $\\le1.08$ & $\\le36$\\\\ 110 & 50 & 0.011 & $\\le3$ & $\\le1.18$ & $\\le30$\\\\ \\hline \\eVS 22 & 10 & 0.011 & $\\le1$ & $\\le1.08$ & $\\le180$\\\\ 22 & 10 & 0.011 & $\\le2$ & $\\le1.29$ & $\\le127$\\\\ 22 & 10 & 0.011 & $\\le3$ & $\\le1.48$ & $\\le\\phantom{1}95$\\\\ \\hline 4.4 & 2 & 0.011 & $\\le1$ & $\\le1.36$ & $\\le570$\\\\ 4.4 & 2 & 0.011 & $\\le2$ & $\\le1.69$ & $\\le368$\\\\ 4.4 & 2 & 0.011 & $\\le3$ & $\\le1.97$ & $\\le271$\\\\ \\hline \\end{tabular}\\\\ {\\bfseries Note.} \\mbox{The binary separation $a=41\\,R_\\odot$ corresponds to $x_a=a/R_\\star=2.4$.} \\end{table} With the mean molecular weight per H atom, $\\mu\\approx1.4$,\\footnote{% The helium abundance per H atom is $\\approx$10\\% \\citep{Wilms2000}. \\label{footnote:HeAbund}} the~continuity equation $\\dot M_\\star \\;=\\; \\mu\\, m_\\mathrm{H}\\, n_\\mathrm{H}(r) \\:\\cdot\\: 4\\pi r^2\\, v(r)$ gives the following estimate for the hydrogen density profile: \\begin{equation} n_\\mathrm{H}(r) \\;=\\; \\frac{\\dot M_\\star/(\\mu\\,m_\\mathrm{H})}{4\\pi R_\\star^2 v_\\infty} \\:\\cdot d(r/R_\\star) \\;\\;\\mathrm{with}\\;\\; d(x) \\;=\\; x^{-2}/f(x) \\label{eq:nr} \\end{equation} Using the parameters of HDE\\,226868 ($\\dot M_\\star=3\\!\\times\\!10^{-6}\\,M_\\odot\\,\\mathrm{yr}^{-1}$, $R_\\star=17\\,R_\\odot$; \\citealp{Herrero1995}; also summarized by \\citealt{Nowak1999_II}, Table~1), \\Eq{eq:nr} predicts \\begin{equation} n_\\mathrm{H}(r) \\;\\;=\\;\\; 2.2\\!\\times\\!10^{10}\\,\\mathrm{cm}^{-3} \\;\\cdot\\; d(r/R_\\star) \\;\\;. \\label{eq:nr_numbers} \\end{equation} \\Eq{eq:nr} can only be solved within the model of \\Eq{eq:f}% \\mbox{\\ifApJ{---}{ --~}}or~any other model for $f$ in \\Eq{eq:vr} with $f(x)\\ge f_0$\\ifApJ{---}{ -- }for \\begin{equation} d \\;\\;<\\;\\; f_0^{-1} \\quad\\quad\\mbox{and}\\quad\\quad 1 \\;\\;\\le\\;\\; x \\;\\;\\le\\;\\; \\sqrt{1/(f_0d)\\;} \\;\\;. \\label{eq:density_solution} \\end{equation} For $f_0\\approx0.01$, the value $d\\approx190$\\ifApJ{---}{ -- }% which would be required to explain the density\\footnote{% A plasma of fully ionized H and He contains $\\approx$1.2 electrons per H atom. } $n_\\mathrm{H,1} \\approx 4.2\\!\\times\\!10^{12}\\,$cm$^{-3}$ obtained from the unshifted Mg\\,\\textsc{xi} triplet (\\Sect{sec:HeTriplets})% \\ifApJ{---}{ -- } can never be reached within our model for the continuous spherical wind. This result shows that the density is likely to be overestimated by an $R$-ratio analysis which ignores the strong UV-flux of the O9.7 star. Table~\\ref{tab:density_solutions} lists therefore some solutions to \\Eq{eq:nr_numbers} for much lower densities as well. Due to the additional constraint that the radial velocity is less than 100\\,km\\,s$^{-1}$ (Table~\\ref{tab:MgXIlines}), we suggest that the first emission component stems from close to the stellar surface. Although the simple wind model of \\Eq{eq:f} may not be appropriate in this region, the results of Table~\\ref{tab:density_solutions} are rather insensitive to the assumptions on the velocity law, as a wide range of $\\beta$ values was considered, but mostly depend on the wind's initial velocity $v_0$, for which we have used a reasonable estimate. Note that $f_0$ and $d$ also depend on $v_\\infty$, but their product, which is important in \\Eq{eq:density_solution}, does not. In spite of the systematic errors of the (absolute) density analysis with the $R$-ratio, we infer that the second plasma component is much denser relative to the first one. As it is seen at a larger redshift of 400--1000\\,km\\,s$^{-1}$, we favor its identification with the focused wind. The two emission components could also be caused coincidentally by dense clumps in the stellar wind \\citep[which are common for O-stars; e.g.,][]{Oskinova2007,LepineMoffat2008}, but the interpretation as a slow base of the wind close to the stellar surface and a focused wind between the accreting black hole and its donor star provides a consistent description of both emission components: an undisturbed wind (with $v_\\infty$ and $f_0$ as above, and $\\beta=1.05\\rightarrow1.6$) would reach a velocity of $v(a)\\approx900\\leftarrow1200$\\,km\\,s$^{-1}$ in the distance of the black hole. The focused wind in the orbital plane would then be detected with a projected velocity $v(a)\\sin i=500\\leftarrow700$\\,km\\,s$^{-1}$ at $\\phi_\\mathrm{orb}=0$. While photoionization reduces the efficiency of acceleration for the spherical wind, the denser focused wind is less strongly affected due to self-shielding and reaches the expected velocity. It is an observational fact that the focused wind has another ionization structure: optical emission lines (H$\\alpha$, He\\,\\textsc{ii} $\\lambda4686$) from ions which only exist at a low ionization parameter have been observed in the focused stream \\citep[see, e.g.,][]{GiesBolton1986_III,Ninkov1987b,Gies2003}. \\mbox{For the spherical wind,} however, \\citet{Gies2008} have conjectured that the part between \\Cyg{} and the donor star might be completely ionized in the soft state. Our observations show that the situation in the hard state may be similar. \\subsection{Modeling of the Photoionization Zone} \\label{sec:XSTAR} We use the photoionization code \\textsc{XSTAR}~2.1ln7b\\footnote{% See \\url{http://heasarc.gsfc.nasa.gov/lheasoft/xstar/xstar.html}\\ifApJ{}{.}} \\citep{Kallman2001} to model the photoionization zone. As the latter is quite complex (due to the inhomogeneous wind density, which is strongly entangled with the X-ray flux), we do not claim to describe the photoionized wind self-consistently, but only want to derive a first approximation. For an optically thin plasma, the relative population of a given atom's ions is merely a function of the ionization parameters \\begin{equation} \\xi(r) \\;\\;=\\;\\; \\frac{L}{n_\\mathrm{H}\\,r^2} \\;\\;=\\;\\; \\frac{L_{37}}{n_{13}\\,r_{12}^2} \\;\\;\\mathrm{erg}\\;\\mathrm{cm}\\;\\mathrm{s}^{-1} \\;\\;, \\label{eq:ionizationParameter} \\end{equation} where $L_{37}$ is the ionizing source luminosity above 13.6\\,eV in~$10^{37}\\,\\mathrm{erg}\\;\\mathrm{s}^{-1}$, $n_{13}$ is the hydrogen density in $10^{13}\\,\\mathrm{cm}^{-3}$ and $r_{12}$~is the distance from the source in $10^{12}\\,\\mathrm{cm}$. \\begin{figure}\\centering \\includegraphics[width=0.95\\columnwidth]{f13_color.eps} \\caption{Relative population of the ionization stages of iron as a function of the ionization parameter $\\xi=L/(n_\\mathrm{H}\\,r^2)$, as calculated with \\textsc{XSTAR} for $L_{37}=3.5$ and $n_{13}=0.01$.} \\label{fig:IonizationBalance} \\end{figure} \\begin{table}{\\centering \\ifApJ{}{\\vskip-5mm} \\caption{Ionization Parameters for Peak Ion Populations} \\label{tab:ionParameter} \\begin{tabular}{cccc} \\hline \\hline Ion & log$\\:\\xi$ & FWHM(log$\\:\\xi$) & \\eVS $r_{12}=\\left(L_{37}/n_{13}/10^{\\log\\xi}\\right)^{1/2}$\\\\ \\hline Fe\\,\\textsc{xxvi} & 2.91 & 0.27 & 0.6--0.8 \\\\ Fe\\,\\textsc{xxv} & 2.75 & 0.32 & 0.7--1.0 \\\\ Fe\\,\\textsc{xxiv} & 2.65 & 0.30 & 0.8--1.1 \\\\ Fe\\,\\textsc{xxiii} & 2.58 & 0.28 & 0.8--1.1 \\\\ Fe\\,\\textsc{xxii} & 2.51 & 0.31 & 0.9--1.3 \\\\ Fe\\,\\textsc{xxi} & 2.44 & 0.36 & 0.9--1.4 \\\\ Fe\\,\\textsc{xx} & 2.32 & 0.44 & 1.0--1.7 \\\\ Fe\\,\\textsc{xix} & 2.12 & 0.40 & 1.2--1.9 \\\\ Fe\\,\\textsc{xviii} & 2.01 & 0.23 & 1.6--2.0 \\\\ Fe\\,\\textsc{xvii} & 1.97 & 0.14 & 1.8--2.1 \\\\ \\hline Ca\\,\\textsc{xx} & 2.69 & 0.56 & 0.7--1.2 \\\\ Ca\\,\\textsc{xix} & 2.39 & 0.54 & 0.9--1.6 \\\\ \\hline Ar\\,\\textsc{xviii} & 2.55 & 0.65 & 0.7--1.5 \\\\ Ar\\,\\textsc{xvii} & 2.22 & 0.50 & 1.0--1.8 \\\\ \\hline S\\,\\textsc{xvi} & 2.38 & 0.69 & 0.8--1.7 \\\\ S\\,\\textsc{xv} & 2.08 & 0.43 & 1.2--2.0 \\\\ \\hline Si\\,\\textsc{xiv} & 2.21 & 0.65 & 0.9--2.0 \\\\ Si\\,\\textsc{xiii} & 1.96 & 0.39 & 1.5--2.4 \\\\ Si\\,\\textsc{iv} & 1.09 & 0.18 & 4.7--5.7 \\\\ \\hline Mg\\,\\textsc{xii} & 2.06 & 0.58 & 1.2--2.3 \\\\ Mg\\,\\textsc{xi} & 1.82 & 0.31 & 1.8--2.6 \\\\ \\hline Ne\\,\\textsc{x} & 1.94 & 0.40 & 1.6--2.5 \\\\ Ne\\,\\textsc{ix} & 1.74 & 0.24 & 2.1--2.8 \\\\ \\hline O\\,\\textsc{viii} & 1.80 & 0.26 & 1.9--2.6 \\\\ O\\,\\textsc{vii} & 1.62 & 0.22 & 2.5--3.2 \\\\ \\hline N\\,\\textsc{v} & 1.51 & 0.01 & 3.25--3.30 \\\\ \\hline C\\,\\textsc{iv} & 1.48 & 0.19 & 3.3--4.1 \\\\ \\hline \\end{tabular}\\\\} ~\\\\ {\\bfseries Notes.} We list the range in ionization parameter and distance for the maximum ion population (\\Fig{fig:IonizationBalance}) obtained in an \\textsc{XSTAR} simulation with $L_{37}=3.5$ and $n_{13}=0.01$. Note that the binary separation in units of $10^{12}\\,$cm is $a_{12}=2.9$. \\end{table} For \\Cyg{} as observed in the hard state, extrapolation of the unabsorbed model obtained by our analysis (Table~\\ref{tab:continuum}) gives $L_{37}\\approx3.5$. Although there are obviously strong variations in the density, the \\textsc{XSTAR} calculation has been performed with constant $n_{13}=0.01$, which is the average of $n_\\mathrm{H}(r)$ from \\Eq{eq:nr_numbers} for $R_\\star\\le r\\le a$. \\Fig{fig:IonizationBalance} shows the resulting population of Fe ions. The ionization parameters at which the population distributions of some ions of interest peak, as well as the corresponding full width at half maximum (FWHM), are presented in Table~\\ref{tab:ionParameter}. We have also included Si\\,\\textsc{iv}, N\\,\\textsc{v}, and C\\,\\textsc{iv} for a comparison with the work of \\citet{Vrtilek2008} and \\citet{Gies2008}. The H- and He-like ions detected with \\Chandra{} only appear at considerable distance from the X-ray source. Taking into account that the photons emanating from \\Cyg{} first propagate through a lower density wind and that the (eventually stalled) wind of higher density is only reached in the vicinity of the star, the actual distances will be even larger than the $r_{12}$ values quoted in Table~\\ref{tab:ionParameter}. \\subsection{Origin of Redshifted X-Ray Absorption Lines} \\begin{figure}\\centering \\includegraphics[width=0.95\\columnwidth]{f14.eps} \\caption{Gray-scale image of $v_\\mathrm{rad,BH}(\\vecr)$ of \\Eq{eq:proj_velocity}, i.e., projected wind velocity against the black hole. A spherically symmetric velocity law given by \\Eqs{eq:vr} and (\\ref{eq:f}) with $v_\\infty=2100\\,\\mathrm{km\\,s}^{-1}$, $f_0=0.01$, and $\\beta=1.05$ is assumed. The star and the black hole are shown as filled black circles; the size of the latter is the accretion radius $2GM_\\mathrm{BH}/v(a)^2$. On the black circle passing through the center of the star and through the black hole $v_\\mathrm{rad,BH}(\\vecr)$ is $0$ since $\\alpha(\\vecr)=90^\\circ$. \\emph{\\mbox{Positive} velocities} (+) are seen only \\emph{within this circle}; a lighter gray means a larger redshift. \\emph{Negative velocities} ($-$) can likewise only occur \\emph{outside of this circle}; a lighter gray here means a larger blueshift. \\ifApJ{Lines of {\\color{red}the}}{The dashed gray lines of} constant \\ifApJ{$v$}{$v_\\mathrm{rad,BH}$} are shown from $-1400\\,$km\\,s$^{-1}$ to $+1000$\\,km\\,s$^{-1}$ in steps of 200\\,km\\,s$^{-1}$. The two highlighted sectors contain all observable lines of sight toward the black hole (after rotation in this plane) for the inclination $i=35^\\circ$. The labels show the corresponding orbital phases. } \\label{fig:proj_velocity} \\end{figure} Additional constraints on the accretion flow can be derived by considering the Doppler shifts observed in absorption lines (Sections~\\ref{sec:absorptionLines} and \\ref{sec:absorptionLineSeries}). Models for the wind velocity like that of \\Eqs{eq:vr} and (\\ref{eq:f}) predict a velocity $\\vecv(\\vecr)$ at the position $\\vecr$ in the stellar wind. But only the projection of $\\vecv$ against the black hole, $v_\\mathrm{rad,BH}$, can eventually be observed as radial velocity in an X-ray absorption line. With the angle $\\alpha(\\vecr)$ between $\\vecv(\\vecr)$ and the direction from $\\vecr$ toward the black hole, $v_\\mathrm{rad,BH}(\\vecr)$ is \\begin{equation} v_\\mathrm{rad,BH}(\\vecr) \\;\\;=\\;\\; \\cos\\alpha(\\vecr) \\;\\cdot\\; |\\vecv(\\vecr)| \\;\\;. \\label{eq:proj_velocity} \\end{equation} Assuming a velocity field with radial symmetry with respect to the star, $\\alpha(\\vecr)$ is just the angle at $\\vecr$ between the star and the black hole. For example, $\\alpha$ is $90^\\circ$ (and $v_\\mathrm{rad,BH}$ is therefore 0) on the sphere containing the center of the star and the black hole diametrically opposed. Redshifted absorption lines can only be observed from wind material inside this sphere, while the part of the wind outside of it is always seen at a blueshift -- a fact which is independent of the assumed velocity field as long as it is directed radially away from the star. For a spherically symmetrical wind model, any line of sight can be rotated to an equivalent one in a half-plane limited by the binary axis. Only a sector with half-opening angle~$i$ can be observed unless the system's inclination is $i=90^\\circ$. \\Fig{fig:proj_velocity} shows the projected velocity $v_\\mathrm{rad,BH}(\\vecr)$ for a simple wind model and two sectors containing all observable lines of sight toward \\Cyg{} for an inclination of $i=35^\\circ$. We now investigate the region where the projected wind velocity (\\Fig{fig:proj_velocity}) is compatible with the observed Doppler shifts (Tables~\\ref{tab:H-He-like_velocityShifts} and \\ref{tab:lineSeries}). We are confident that this method allows for the identification of the absorption regions, as the low observed velocities are always found close to the $\\alpha=90^\\circ$ sphere -- independent of the wind model. For most of the investigated ions (\\Ion{Ne}{x}, \\Ion{Mg}{xii}, \\Ion{Si}{xiv} and \\textsc{xiii}, \\Ion{S}{xv}, \\Ion{Ar}{xvii}, \\Ion{Fe}{xix} and \\textsc{xviii}) the inferred distance from the black hole agrees with the predictions of Table~\\ref{tab:ionParameter}, e.g., the projected velocity $-117\\,$km\\,s$^{-1}\\le v_\\mathrm{Ne\\,X}\\le -69\\,$km\\,s$^{-1}$ measured for \\Ion{Ne}{x} is (during $0.93\\le\\phi_\\mathrm{orb}\\le1$) obtained at $r_{12}=1.78\\pm0.07$. For many other ions, both results are still very similar: e.g., \\Ion{Ne}{ix} with $-214\\,$km\\,s$^{-1}\\le v_\\mathrm{Ne\\,IX}\\le -123\\,$km\\,s$^{-1}$ is expected at $r_{12}=1.95\\pm0.07$ from the $v_\\mathrm{rad,BH}$ model and at $r_{12}=2.1$--2.8 from the photoionization model. Only for the highly ionized iron lines, the small distances are -- as already anticipated in \\Sect{sec:XSTAR} -- underestimated by the \\textsc{XSTAR} simulation run with constant average density, which overestimates the wind density close to the black hole. For example, the velocity range $-9\\,$km\\,s$^{-1}\\le v_\\mathrm{Fe\\,XXIV}\\le52\\,$km\\,s$^{-1}$ measured for \\Ion{Fe}{xxiv} corresponds to $r_{12}=1.51\\pm0.07$, while the population of this ion peaks at $r_{12}=0.8$--1.1 within the model presented in Table~\\ref{tab:ionParameter}. \\subsection{Previous \\Chandra{} Observations}\\label{sec:prevObs} \\begin{figure}\\centering \\includegraphics[width=0.95\\columnwidth]{f15.eps} \\caption{Orbital phase and mean ASM count rate of the previously reported \\Chandra{} observations of \\Cyg{} (ObsID~3814 is presented in this paper). } \\label{fig:observations} \\end{figure} In previous \\Chandra-HETGS observations of \\Cyg, the stellar wind was seen under different viewing angles, or the X-ray source was in other spectral states (see \\Fig{fig:observations}), which probably means that the properties of the wind were also different. \\citet{Schulz2002} have analyzed the 14\\,ks \\ifApJ{}{observation }ObsID~107 performed in 1999 October (at $\\phi_\\mathrm{orb}\\approx0.74$),\\footnote{% Note that \\citet{Schulz2002} and \\citet{Miller2005} quote an erroneous date and thus the wrong orbital phase $\\phi_\\mathrm{orb}=0.93\\ne0.74$ for this observation.} when \\Cyg{} was in a transitional state. They derive a neutral column density of $N_\\mathrm{H}=6.2\\!\\times\\!10^{21}\\:\\mathrm{cm}^{-2}$ from prominent absorption edges (see also Table~\\ref{tab:neutralAbundances}) and detect some emission and absorption lines with indications of P~Cygni profiles. \\citet{Miller2005} have investigated the focused wind with the 32\\,ks \\ifApJ{}{observation }ObsID~2415 of \\Cyg{} in an intermediate state, performed in 2001 January (at $\\phi_\\mathrm{orb}\\approx0.77$). They report absorption and emission lines of H- and He-like resonance lines of Ne, Na, Mg, and Si with a mean redshift of $\\approx$100\\,km\\,s$^{-1}$, as well as some lines of highly ionized Fe and Ni. The column density is $6.2\\!\\times\\!10^{21}\\:\\mathrm{cm}^{-2}$ \\citep{Miller2002}. \\citet{Marshall2001} describe Ly\\,$\\alpha$ and He\\,$\\alpha$ absorption lines, redshifted by $(450\\pm150)\\:\\mathrm{km\\,s}^{-1}$ from the 12.6\\,ks \\ifApJ{}{observation }ObsID~1511 of \\Cyg{} in the hard state, performed in 2001 January (at $\\phi_\\mathrm{orb}\\approx0.83$). \\citet{ChangCui2007} report dramatic variability in the 30\\,ks \\ifApJ{}{observation }ObsID~3407 performed in 2001 October (at $\\phi_\\mathrm{orb}\\approx0.88$), when \\Cyg{} reached its soft state. While a large number of absorption lines (mostly redshifted, but not by a consistent velocity) is identified in the first part of their observation, most of them weaken significantly or cannot be detected at all during the second part. The complete ionization of the wind due to a sudden density decrement is given as a possible explanation. \\citet{Feng2003} detect asymmetric absorption lines with the 26\\,ks \\ifApJ{}{observation }ObsID~3724 of \\Cyg{} in the soft state, performed in 2002 July (at $\\phi_\\mathrm{orb}=0$). The line centers are almost at their rest wavelengths, but the red wings are more extended, especially for the transitions of highest ionized ions, which they explain by the inflowing focused wind reaching both the highest redshift and ionization parameter closest to the black hole. The interpretation of the absorption lines presented in the previous section describes our observation consistent with wind and photoionization models. It can, however, not be applied to all of the other observations; the model of \\Fig{fig:proj_velocity} neither predicts a conspicuously high redshift at $\\phi_\\mathrm{orb}\\approx0.83$ and at a higher soft X-ray flux (ObsID~1511), nor a positive redshift at $\\phi_\\mathrm{orb}\\approx0.77$ at a still higher flux (ObsID~2415, see \\Fig{fig:observations}). Inhomogeneities (e.g., density enhancements in shielding clumps) and asymmetries of the wind (e.g., due to noninertial forces in the binary system) may therefore play an important role in these cases. \\subsection{Summary} In this paper, we have presented a \\Chandra-HETGS observation of \\Cyg{} during superior conjunction of the black hole, which allows us to detect the X-ray absorption signatures of the stellar wind of HDE\\,226868. The light curve near $\\phi_\\mathrm{orb}=0$ is shaped by absorption dips; these are, however, excluded from this analysis by selecting nondip times of high count rate only. At a flux of $\\sim$0.25\\,Crab, we have to deal with moderate pile-up in the grating spectra, for which can be accounted very well with the \\texttt{simple\\_gpile2} model. The continuum of both \\Chandra's soft and \\RXTE's broadband X-ray spectrum has been described consistently by a single model, consisting of an empirical broken power-law spectrum with high-energy cutoff (which is typical for the hard state) and a subordinated disk component with an inner radius close to the ISCO. The joint modeling of both spectra reveals the presence of both a narrow and a relativistically broadened fluorescence Fe K$\\alpha$ line. \\Chandra{} has resolved absorption edges of neutral atoms with an overabundance of metals, suggesting an origin not only in the ISM, but also in the stellar wind of the evolved supergiant. The previously suspected anticorrelation of neutral column densities and X-ray flux, which is due to photoionization, is confirmed. Absorption lines of highly ionized ions are produced where the wind becomes extremely photoionized. For H- and He-like ions, Lyman and He-series are detected up to the Ly or He$\\:\\epsilon$ line, which we use to measure the column density and velocity of absorbing ions. For the wealth of Fe L-shell transitions, column densities can best be obtained by directly using a physical model for the complete line series of an ion. The nondip spectrum shows almost no Doppler shifts, probably indicating that we have excluded the focused wind by our selection of nondip times. We have also detected two plasma components in emission by two pairs of i- and f-lines of He-like Mg\\,\\textsc{xi}. The first one is roughly at rest and we identify it with the base of the spherical wind close to the stellar surface. The second plasma component is denser and observed at a redshift that is compatible with the focused wind. A simple \\textsc{XSTAR} simulation indicates that most of the observed ions only exist in a distance of the black hole, where the velocity of a spherical wind, projected against the black hole, is low -- which is a consistent explanation of the small Doppler shifts observed in absorption lines. We review the previously reported \\Chandra{} observations of \\Cyg{} and find that not all of them can be described in the same picture, as the wind may be affected by asymmetries or inhomogeneities. A detailed spectroscopy of the absorption dips, which might shed more light on the focused wind, will be presented in a subsequent paper." }, "0808/0808.1890_arXiv.txt": { "abstract": "The orbital parameters of extra-solar planets have a significant impact on the probability that the planet will transit the host star. This was recently demonstrated by the transit detection of HD 17156b whose favourable eccentricity and argument of periastron dramatically increased its transit likelihood. We present a study which provides a quantitative analysis of how these two orbital parameters affect the geometric transit probability as a function of period. Further, we apply these results to known radial velocity planets and show that there are unexpectedly high transit probabilities for planets at relatively long periods. For a photometric monitoring campaign which aims to determine if the planet indeed transits, we calculate the expected transiting planet yield and the significance of a potential null result, as well as the subsequent constraints that may be applied to orbital parameters. ", "introduction": "With the number of known extra-solar planets exceeding 300, statistical interpretations of the distribution of orbital parameters are becoming increasingly significant. These parameter distributions help us unlock the mysteries surrounding the planet formation process to which many challanges have been presented, not the least of which contains the mechanisms that drive planetary migration \\citep{arm07}. \\citet*{for08b} showed that transit light curves in particular can be used to characterize orbital eccentricities and hence give further insight into the global eccentricity distribution. In terms of the sheer number of transit light curves, the major contributors have been the shallow wide-field surveys such as the Transatlantic Exoplanet Survey (TrES) \\citep{man07}, the XO project \\citep{joh08}, the Hungarian Automated Telescope Network (HATNet) \\citep{pal08}, and SuperWASP \\citep{and08}. In addition, there have been at least five cases in which planetary transits were detected through photometric follow-up of planets already known via their radial velocity (RV) discoveries. These five planets are HD 209458b \\citep{cha00,hen00}, HD 149026b \\citep{sat05}, HD 189733b \\citep{bou05}, GJ 436b \\citep{gil07}, and HD 17156b \\citep{bar07a}. The case of HD 17156b is of particular interest since it is a 21.2 day period planet which happens to have a large eccentricity ($e = 0.67$) and an argument of periastron which places the periapsis of its orbit in the direction toward the observer and close to parallel to the line of sight, resulting in an increased transit probability. Conversely, the dominant sources of RV planet discoveries have been the California \\& Carnegie Planet Search \\citep{mar97} and the High Accuracy Radial velocity Planet Searcher (HARPS) \\citep{pep04} teams. However, in the near future we can expect to see larger-scale surveys \\citep*{kan07b} and new instruments \\citep{li08} which will increase both the number and diversity of known planets. There have been suggestions regarding the strategy for photometric follow-up of these radial velocity planets at predicted transit times \\citep{kan07a} and the instruments that could be used for such surveys \\citep{lop06a}. Some attempts have been made to detect these possible transits \\citep{lop06b,sha06} which have thus far been unsuccessful. This paper discusses the effect of orbital parameters on the geometric transit probability of planets. We calculate orbital constraints that may be applied, particularly in the absence of transit signatures in photometric follow-up observations. Section 2 describes how the eccentricity and argument of periastron of known planetary orbits affect transit probability. It further presents applications of this effect to known RV planets and discusses how uncertainties in the orbital parameter values affect the reliability of the ephemeris calculations. In Section 3, we show how orbital constraints can be applied in the absence of a photometrically detected transit signal, and we discuss the potential transit yield and statistical significance of a scenario in which no transits are found in a large sample of RV planets. We summarize and conclude in Section 4. ", "conclusions": "It is still uncertain at this stage how many of the known radial velocity planets transit their parent stars. What is clear is that the eccentricity distribution of the known exoplanets will increase the transit likelihood, making detections for long-period planets, such as HD 17156b, feasible. We have shown in this paper that there is enough potential amongst longer period planets for transit detections to motivate a photometric monitoring campaign at the predicted times of transit for these targets. \\citet{fle08} have shown that long-period transiting planets may yet be discovered through ground-based transit surveys, particularly if data sets from different surveys are combined. As pointed out by \\citet{bar07b}, eccentric planets that have a periastron oriented away from the observer are far more likely to exhibit a secondary than a primary eclipse. The detection of such a secondary eclipse is considerably more challenging than for a primary eclipse since it relies on a minimum level of planetary flux and is best pursued at infrared wavelengths. The discussion in \\S \\ref{i_and_e} shows that even an assumption of $\\omega = 3\\pi/2$ can place constraints on the orbital inclination. A prime candidate for such a study is HD 80606b \\citep{nae01} which has a period of 111.87 days and an eccentricity of 0.927. Scaling Figure \\ref{fig1} to this period and eccentricity yields a secondary transit probability of $\\sim 15$\\%. Many of the results presented in this paper can easily be applied to any system since the results generally scale linearly with the sum of the stellar and planetary radii. Through applying these results to current and future radial velocity planet discoveries, one can choose targets for an efficient observing campaign which may help to discover long-period transiting planets and hence add invaluable information to planetary structure and formation theories." }, "0808/0808.4151_arXiv.txt": { "abstract": "We consider the prospects for testing the dark matter interpretation of the DAMA/LIBRA signal with the Super-Kamiokande experiment. The DAMA/LIBRA signal favors dark matter with low mass and high scattering cross section. We show that these characteristics imply that the scattering cross section that enters the DAMA/LIBRA event rate determines the annihilation rate probed by Super-Kamiokande. Current limits from Super-Kamiokande through-going events do not test the DAMA/LIBRA favored region. We show, however, that upcoming analyses including fully-contained events with sensitivity to dark matter masses from 5 to 10 GeV may corroborate the DAMA/LIBRA signal. We conclude by considering three specific dark matter candidates, neutralinos, WIMPless dark matter, and mirror dark matter, which illustrate the various model-dependent assumptions entering our analysis. ", "introduction": "The DAMA/LIBRA experiment has seen, with $8.2\\sigma$ significance~\\cite{Bernabei:2008yi}, an annual modulation~\\cite{Drukier:1986tm} in the rate of scattering events, which could be consistent with dark matter-nucleon scattering. Much of the region of dark matter parameter space that is favored by DAMA is excluded by null results from other direct detection experiments, including CRESST~\\cite{Angloher:2002in}, CDMS~\\cite{Akerib:2005kh}, XENON10~\\cite{Angle:2007uj}, TEXONO~\\cite{Lin:2007ka,Avignone:2008xc}, and CoGeNT~\\cite{Aalseth:2008rx}. On the other hand, astrophysical uncertainties~\\cite{Brhlik:1999tt,Gondolo:2005hh} and detector effects~\\cite{Bernabei:2007hw} act to open up regions that may simultaneously accommodate the results from DAMA and these other experiments. Following DAMA's latest results, several recent studies~\\cite{Feng:2008dz,Petriello:2008jj,Chang:2008xa,% Fairbairn:2008gz,Savage:2008er} have studied the consistency of DAMA with other direct detection experiments, with varying assumptions and varying conclusions. What is clear, however, is that if DAMA is seeing dark matter, the preferred region of parameter space has dark matter mass in the range $m_X \\sim 1-10~\\gev$ and spin-independent proton scattering cross section $\\sigmaSI \\sim 10^{-5} - 10^{-2}~\\pb$. Although neutralinos have been proposed as a possible explanation~\\cite{Bottino:2003iu}, such low masses and high cross sections are not typical of weakly-interacting massive particles (WIMPs), and alternative dark matter candidates have been suggested to explain the DAMA signal~\\cite{Smith:2001hy,Feng:2008ya,Feng:2008mu,% Foot:2008nw,Feng:2008dz,Khlopov:2008ty,Andreas:2008xy,Dudas:2008eq}. The current state of affairs also makes it abundantly clear that data from complementary experiments is likely required to sort out the true nature of this result and is certainly required to establish definitively the detection of dark matter. Other direct detection experiments may play this role. In this paper, we note that corroborating evidence may come from a very different source, namely, from the indirect detection of dark matter at Super-Kamiokande (Super-K). In contrast to direct detection experiments, which rapidly lose sensitivity at low masses, given physical limitations on threshold energies, Super-K's limits remain strong for low masses. Super-K is therefore poised as one of the most promising experiments to either corroborate or exclude many dark matter interpretations of the DAMA/LIBRA data. In \\secref{relating}, we show that, with a few well-motivated theoretical assumptions, the DAMA and Super-K event rates may be related. Currently published Super-K results do not challenge the DAMA preferred region. In \\secref{projection}, however, we show that there is significant potential for Super-K to extend its reach to dark matter masses from 5 to 20 GeV and provide sensitivity that is competitive with, or possibly much better than, direct detection experiments. In \\secref{models}, we apply our analysis to three specific dark matter candidates that have been proposed to explain DAMA: neutralinos~\\cite{Bottino:2003iu}, WIMPless dark matter~\\cite{Feng:2008ya,Feng:2008dz,Feng:2008mu}, and mirror dark matter~\\cite{Foot:2008nw}. These candidates illustrate and clarify the assumptions entering the analysis. We present our conclusions in \\secref{summary}. As this work was in preparation, a study appeared that also considered testing the dark matter interpretation of DAMA with data from Super-K~\\cite{Hooper:2008cf}. That work focused primarily on a model-independent approach and present Super-K data, considering neutralinos briefly as a case example, whereas this work considers neutralino, WIMPless, and mirror dark matter and emphasizes the much brighter prospects for future Super-K results. ", "conclusions": "\\label{sec:summary} The DAMA/LIBRA signal is currently of great interest, and alternative methods for corroborating or excluding a dark matter interpretation are desired. In this study, we have shown that the preferred DAMA region implies that Super-K, through its search for dark matter annihilation to neutrinos, has promising prospects for testing DAMA. We have given a conservative estimate of the projected sensitivity of Super-K. By using fully contained events, we expect that current super-K bounds may be extended to dark matter masses of 5 to 10 GeV. In the region of most interest for the DAMA result with $m_X \\sim 5-10~\\gev$, the neutralino models of Ref.~\\cite{Bottino:2003iu} and WIMPless models can potentially be tested, provided the sensitivities expected at this low mass range are actually realized. For mirror matter, however, the mass range of interest is $10-30~\\gev$, and it is unlikely that Super-K can place limits on this model. We thus have the intriguing prospect that the direct detection result of DAMA/LIBRA could be sharply tested by an indirect detection experiment in the very near future." }, "0808/0808.2754.txt": { "abstract": " ", "introduction": "\\label{intro} This is a scientific strategy for the detection and characterization of extrasolar planets; that is, planets orbiting other stars. As such, it maps out--over a 15-year horizon--the techniques and capabilities required to detect and measure the properties of planets as small as Earth around stars as large as our own Sun. It shows how the technology pieces and their development fit together to achieve the strategy's primary goal: if planets like Earth exist around stars within some tens of light years of our own Solar System, those planets will be found and their basic properties characterized. Essential to this strategy is not only the search for and examination of individual planets, but also a knowledge of the arrangement, or \\emph{architecture}, of planetary systems around as large a number of stars as possible; this is the second goal of the strategy. The final goal of the strategy is the study of disks around stars, important both to understand the implications of the variety of exoplanet systems for planet formation, and to determine how many nearby stars have environments around them clean enough of debris that planets may be sought and, if found, characterized. It is important that the program target planets as small as the Earth, if we are to determine whether the conditions we find on our own world are a common outcome of planetary evolution. The only example of a habitable world we have is our own one-Earth-mass planet, and indeed our nearest neighbor, Venus, is nearly the same mass but uninhabitable by virtue of closer proximity to the Sun. Searching for planets, for example, five times the mass of Earth is easier but should they turn out to lack habitable atmospheres we would not know whether this is by chance or a systematic effect of the higher mass. This goal by no means excludes detection and study of larger planets, which themselves will provide a wealth of discoveries and new knowledge, but the Task Force strongly believes that public interest ultimately lies in knowing whether planets like our own are a common outcome of cosmic evolution. For that reason, we have set the bar high. The charge to the Exoplanet Task Force (``ExoPTF\"), which conducted the study for the Astronomy and Astrophysics Advisors Committee (``AAAC\"), is reproduced after the bibliography. The ExoPTF membership, listed above, was chosen from the astronomical commu\"nity to represent the range of techniques and expertise involved in exoplanet research today: in instrumentation, observation, and theory. The Task Force met five times from February to September 2007. Representatives of NASA and the NSF were present as \\emph{ex officio} members, and a number of presentations from outside experts were solicited to provide the Task Force with the most up to date information available. Early in the process, ``white papers\" were solicited on the Task Force's public website, inviting interested individuals to submit ideas for techniques, missions, and theoretical investigations in the area of planet search and detection. Eighty-five white papers were received, read, and discussed by the Task Force. A number of the white papers provided important information that, when verified with outside experts, influenced the ultimate strategy; the sum total of the white papers provided an impressive indication of the breadth and depth of thinking in the astronomical community. Periodic briefings on the Task Force's progress were provided to the AAAC committee, and general reports were provided to the community at selected scientific conferences and workshops. A draft report was reviewed by the AAAC in October 2007 and a final report delivered in March 2008. The principal goal of this strategy--the detection of habitable planets--is perhaps among the most challenging and at the same time most rewarding goals in modern science. In a sense, habitable planets are in an ``anti-sweet spot\", wherein the position they occupy around nearby Sun-like stars does not correspond to the highest senstivity zone for any of the techniques discussed here. Further, Earth-sized planets will have extremely low signal-to-noise, making analysis of the detected signatures of these objects extremely difficult. The Task Force was encouraged, however, by the extraordinary progress made in the last half-decade in detecting planets down to less than ten times the mass of the Earth, in characterizing extrasolar giant planets spectroscopically, in technological progress accomplished in key approaches that will be needed to detect and study Earth-sized planets, and in the innovative and entrepeneurial character of the community in proposing and in some cases executing novel approaches to detection and characterization. It is the vigor of the field and its practitioners that ensures ultimate success in this difficult endeavor. The Task Force firmly believes that the strategy presented herein is reasonably paced, technically possible, and scientifically compelling. It is robust against surprises both in the areas of science ({\\it e.g.,} frequency of occurrence of planets) and technological development. But the strategy requires a level of stamina not typical of the Federal enterprise. Important and exciting results will come from the near-term portions of the strategy (\\emph{i.e.,} within 5 years), but the ultimate goal of finding and characterizing a planet like our own, around a star like our Sun, will require the full 15-year time horizon for its realization. \\pagebreak % ============================================================================ % Begin New PART - SCIENTIFIC AND PHILOSOPHICAL SIGNIFICANCE % ============================================================================ \\chapter{Scientific and Philosophical Significance of Detecting other Earths} The idea that humankind's home in the cosmos is just one of countless worlds has been favored by many cultures throughout human history. The thread of western thought on this matter goes back to antiquity, but the view of the cosmos promulgated in medieval Europe put the Earth in a special place and discouraged speculation on the possibility of other inhabited worlds. Nonetheless, the German philosopher and theologian Albertus Magnus wrote in the 13th century: ``Do there exist many worlds, or is there but a single world? This is one of the most noble and exalted questions in the study of Nature.\" Two centuries later, Nicolas of Cusa asserted on a philosophical basis that the heavens were teeming with inhabited worlds. The notion of a plurality of worlds was put on a scientific footing in the mid-16th century by Copernicus' cosmology that displaced the Earth from the center of the Solar System, so that it became possible to imagine an abundance of unseen worlds in the form of other solar systems, as Giordano Bruno wrote: \\\\ \\begin{quote} \\emph{``There are countless suns and countless earths all rotating around their suns in exactly the same way as the seven planets of our system. We see only the suns because they are the largest bodies and are luminous, but their planets remain invisible to us because they are smaller and non-luminous. The countless worlds in the universe are no worse and no less inhabited than our Earth.\"}\\\\ \\end{quote} These ideas were not popular in the chilly counter-Reformation environment of the times, and indeed Bruno was burned at the stake in 1600. Less than a decade later, however, Galileo Galilei saw through his telescope first that our Moon--rather than being a perfect sphere--was in fact a mountainous world in its own right, and then that planet-like bodies (moons) orbited Jupiter just as our Moon orbits the Earth. The progressive displacement of humanity's position from the cosmic center became irreversible. Today we know the Earth is one of countless bodies that orbit the Sun; the Sun orbits the center of the Milky Way Galaxy along with $\\simeq 200$ billion other stars; there are tens of billions of galaxies like the Milky Way in the cosmos. And yet, one special aspect of the Earth remains: of all the planets in the solar system, only Earth demonstrably is teeming with life. If life exists on other planets or moons of the Solar System, it does not have the profound impact on the properties of those bodies that life has on Earth, and for all we know Earth is the only abode of life around the Sun. No world in the solar system is remotely suitable for life as we know it in the way that Earth is; we must look beyond our own Solar System to the stars if we are to determine whether other habitable worlds like the Earth exist. To complete the Copernican revolution, we must detect and characterize ``extra-solar\" planets the size of our home world. Beginning in the early 1990s, planets around other stars began to be detected, first by their indirect effects on the motion of their parent stars, then directly. Today, over 260 planets are known to exist around other stars, about 10\\% of which can be studied directly. All of these, with just a couple of exceptions, are more than ten times more massive than the Earth, and we could not today detect our own home world in orbit around a Sun-like star a mere 4 light years away--the distance to the closest star Proxima Centauri. But we are close to detecting and characterizing Earth-sized planets around smaller stars, and the means required to do so around a star like the Sun are so well understood that we argue here that the required technologies could be fielded within fifteen years. A strategy that maximizes the chances for success while minimizing risk forms the core of this report. Complementary to detecting and characterizing individual planets the size of the Earth is determining whether the architecture of our own solar system--rocky planets huddled close to the parent star and multiple giant planets in more distant orbits--is typical. We have enough information today to say that upwards of 15\\% of Sun-like stars harbor giant planets, and that planets down to the size of Neptune also seem to be common, but have no information on the occurrence of planets the mass or size of the Earth. Because the formation of rocky planets like Earth may have been quite different from that of giant planets, extrapolation is risky. The scientific importance of understanding both the statistical occurrence of Earth-sized planets and the architectures of planetary systems is high, because these are crucial to understanding the processes by which planetary systems form. They therefore form a part of our strategy in conjunction with direct detection and characterization, as well as the study of planet forming disks and disk debris. Together, these elements will ensure that, no matter what the ultimate commonality or rarity of Earth-sized planets might be, the end result of the strategy will be a deep understanding of how planets fit into the overall scheme of cosmic evolution. Humankind stands today at the threshold of answering one of humankind's most ancient questions: is our home world the only suitable abode for life like us in the cosmos? Most humans would answer no, but only because the Copernican mindset is so firmly a part of our modern culture. What if we were rare enough that no such other Earth existed within the reach of modern astronomical telescopes, on the ground or in space? How would we react to the confirmed emptiness the universe presented to us? Would our specialness lead to a profound sense of faith, or a deep loneliness? Would it stiffen humankind's resolve to survive, to solve its political and environmental problems? Would our solitary state lead us to be more mindful of the precious uniqueness of our intellect and awareness? Conversely, were we to find a planet like the Earth--rocky, similar in size to our world, possessed of an atmosphere-- around nearby stars, our perception of the cosmos around us would change in an equally profound fashion. We would look up still with wonder, but a handful of stars in our night sky will forever after hold a special place in our imagination, tempting us with wild dreams of flight. Surely that too would make us refocus our energies to hasten the day when our descendants might dare to try to bridge the gulf between two inhabited worlds. \\\\ \\begin{quote} {\\emph{``Which one is it, Mommy?\" asked the older of her two children. They had walked away from the campfire, and gazing now at the familiar pattern of stars in the night sky, a question far different from any ever asked by thousands of generations of human beings drifted off in the cool night air. ``Look at the bright star over there,\" the woman responded to them ``now move your eyes a little to the right, and you'll see that slightly fainter star. The planet belongs to that one. It's almost exactly the size of the Earth, is just a little closer to its sun than we are to ours, and the space telescope that your mommy helped build found oxygen in its atmosphere. That world has air that creatures like us could breathe.\" ``Who lives there?\" asked the younger one. ``No one knows,\" the woman replied, ``but maybe they are looking at us, right now, wondering the same thing.\" }} \\end{quote} \\pagebreak % ============================================================================ % Begin New PART - THE KNOWN EXTRASOLAR PLANETS % ============================================================================ \\chapter{The Known Extrasolar Planets} %\\setcounter{section}{0} % - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ", "conclusions": "" }, "0808/0808.3895_arXiv.txt": { "abstract": "We have studied the kinematics traced by the water masers located at the centre of the planetary nebula (PN) K3-35, using data from previous Very Large Array (VLA) observations. An analysis of the spatial distribution and line-of-sight velocities of the maser spots allows us to identify typical patterns of a rotating and expanding ring in the position-velocity diagrams, according to our kinematical model. We find that the distribution of the masers is compatible with tracing a circular ring with a $\\simeq$0\\farcs021 ($\\simeq$100~AU) radius, observed with an inclination angle with respect to the line of sight of 55$\\degr$. We derive expansion and rotation velocities of 1.4 and 3.1 km s$^{-1}$, respectively. The orientation of the ring projected on the plane of the sky, at PA$\\simeq$158$\\degr$, is almost orthogonal to the direction of the innermost region of the jet observed in K3-35, suggesting the presence of a disc or torus that may be related to the collimation of the outflow. ", "introduction": "High angular resolution observations of molecular gas have revealed the presence of dense equatorial discs and tori towards several late asymptotic giant branch (AGB) stars and young planetary nebulae (PNe): for instance, \\citet{bie88}, \\citet{for98}, \\citet{sah98b}, \\citet{buj03}. The interaction between the post-AGB wind and such equatorial structures has been proposed as one of the possible physical mechanisms in shaping the bipolar and multipolar morphologies seen in PNe and proto-PNe (PPNe) \\citep{bal87,mel95,sok00,ick03}. The origin of molecular discs and tori around late AGB stars is not completely clear, but a possible explanation is the presence of a binary system \\citep{mor87,liv88,taa89,sok06}. In this case, when one of the stars enters the AGB phase, some of the ejected material can be retained, generating an extended torus. However, in the case of the high velocity jets observed in some late AGB stars, post-AGB stars, and young PNe \\citep{fei85,sah98a,ima02,rie03} it is believed that a more effective collimation mechanism(s) should be present in the innermost region of the object, such as an accretion disc \\citep{mor87,sok94,sok00} or a stellar magnetic field produced by a rotating star \\citep{pas85,che94,gar97}. Furthermore, recent observations reveal compelling evidence for magnetic fields in post-AGB stars and PNe, associated with equatorial discs and/or jets (e.g. \\citealt{sab07,vle06}). The process for the formation of accretion discs in PPNe has been investigated in detail by \\citet{rey99}. They find that a disc forms when a close binary system (with a substellar companion) undergoes common envelope evolution. For the case in which a low-mass secondary is disrupted during a dynamically unstable mass transfer process, an accretion disc, with a radius of $\\sim$10~AU and a mass of $\\sim$2$\\times10^{-3}$~M$_{\\sun}$, forms within $\\sim$100~yr. Recently, \\citet{rij05} from their 3D simulations based on a two-wind model with a warped disc, suggested that, in order to explain the observed multipolar and point-symmetric shape of PNe, the required discs are quite small ($\\sim$10--100~AU). Furthermore, these disc-like structures should be dense ($10^7-10^8$~cm$^{-3}$) and in Keplerian rotation. Interferometric CO observations show larger toroidal molecular structures with sizes in the range of $\\simeq$1000--6000 AU in PPNe and young PNe. These structures seem to be systematically in expansion with a mean velocity of $\\sim$7~km~s$^{-1}$, such as in M~1-92 \\citep{buj98}, M~2-9 \\citep{zwe97}, \\mbox{M~2-56} \\citep{cas02}, or KjPn~8 \\citep{for98}. These velocities are comparable to or below those found in expanding AGB envelopes \\citep{hug07}. Importantly, rotation has been observed in the Red Rectangle as well as slower expansion ($\\sim$0.8~km~s$^{-1}$), superimposed on rotation, according to \\citet{buj05}. Note that the sizes of the tori are about one or two orders of magnitude more than the sizes of the disc-like structures proposed by \\citet{rey99} and \\citet{rij05}. Maps of water maser emission allow the identification of a disc-like structure with a radius of 0\\farcs12 in IRAS 17347-3139 \\citep{deg04}, which corresponds to $\\simeq$100--750~AU at the source distance \\citep{gom05a}. In this case, the gas kinematics suggests the presence of both rotation and expansion in the disc traced by the water masers. In this context, it is very important to study in detail the kinematics of much smaller disc-like structures that might be related to the collimation of the observed bipolar outflows in some PNe. K3-35 is a young PN that shows a bipolar outflow in optical images \\citep{mir00}. At radio wavelengths, K3-35 exhibits a bright core and two bipolar lobes with an S-shape \\citep{mir01}. The distance to this object has been estimated to be $\\sim$5~kpc \\citep{zha95}, using an statistical method. However, we note the large uncertainty of this type of estimate, since the application of different statistical methods could give distances varying by factors of $\\sim$3 \\citep{phi04}. The characteristic S-shape morphology of the radio lobes can be successfully reproduced by a precessing jet evolving in a dense circumstellar medium \\citep{vel07}. Water maser emission has been found in three regions: two regions located at the tips of the bipolar radio jet about $\\simeq$1$''$ from the centre (regions N and S), and another region towards the core of the nebula within $\\simeq$0\\farcs02 (region C), suggesting the presence of a torus \\citep{mir01}. In addition, OH maser emission has been detected towards the centre of K3-35 (within $\\simeq$0\\farcs04), showing circular polarisation that suggests the presence of a magnetic field \\citep{mir01,gom05b}. We decided to study the spatio-kinematical distribution of the water masers reported by \\citet{mir01} towards the centre of K3-35 in order to identify possible expansion and/or rotation motions. The paper is organised as follows. In Section 2, we present a simple kinematical model of a ring including both expansion and rotation, and we calculate the pattern delineated in the position-velocity diagrams. We also describe the least-squares fit procedure that we used. In Section 3, we present the current observational data of H$_2$O masers in the PN K3-35. We then apply our model to the H$_2$O masers located towards the centre of the PN K3-35, making a comparison between the results and the observations. Finally, in Section 4, we discuss the implications of our results. \\section[]{Rotating and Expanding Ring Model} \\subsection{Model} We assume a narrow, uniform, rotating, and expanding ring of radius $R$, arbitrarily oriented with respect to the line of sight. Its projection on the plane of the sky is an ellipse with semimajor and semiminor axes $a$ and $b$, respectively. We define the two frames of reference shown in Fig.~1. Both coincide with the plane of the sky, one of them has its origin at the centre of the ellipse and is oriented such that the $x'$-axis is along the major axis of the projected ellipse, and the other one has the axes parallel to the RA and Dec axes. The semimajor and semiminor axes are related to the ring radius by $a=R$ and $b=R\\cos i$, where $i$ is the inclination angle between the line of sight and the normal to the ring plane as shown in Fig.~2. The equation of the ellipse is given by \\begin{equation} \\frac{x'^2}{a^2}+\\frac{y'^2}{b^2}=1, \\end{equation} and the transformation equations between the coordinate systems are \\begin{equation} x'=(x-x_0)\\cos\\theta+(y-y_0)\\sin\\theta, \\end{equation} \\begin{equation} y'=-(x-x_0)\\sin\\theta+(y-y_0)\\cos\\theta, \\end{equation} where ($x_0,y_0$) is the position of the centre of the ellipse, and $\\theta$ is the angle between the $x$-axes of the two frames of reference, and is defined as positive clockwise. The angle $\\theta$ is related to the position angle (PA) of the major axis of the ellipse by $\\rmn{PA}=90\\degr-\\theta$. \\begin{figure} \\includegraphics[width=72mm]{fig1.eps} \\caption{Reference systems. Both the $x'y'$- and the $xy$-coordinate systems are in the plane of the sky. The $x'y'$-system has its origin at the centre of the ellipse ($x_0,~y_0$), and $\\theta$ is the angle between the $x$-axis and the $x'$-axis. The $x$ and $y$ axes are parallel to the RA and Dec axes.} \\end{figure} \\begin{figure*} \\includegraphics[width=150mm]{fig2.eps} \\caption{Position-velocity diagrams for a rotating and expanding ring model. The top panels correspond to a positive inclination angle and the bottom panels correspond to a negative inclination angle. The sense of rotation (clockwise or counterclockwise) as seen from the observer is indicated. The $x'$ and $y'$ axes are those defined in Fig.~1.} \\end{figure*} Let $v_s$, $v_{\\rmn{rot}}$, and $v_{\\rmn{exp}}$ be the local standard of rest (LSR) systemic velocity, the rotation velocity, and the expansion velocity of the ring, respectively. Then the observed LSR velocity of a point in the ring can be expressed as \\begin{equation} V_{\\rmn{LSR}}=v_s+\\frac{x'}{a}v_{\\rmn{rot}}\\sin i+\\frac{y'}{a}v_{\\rmn{exp}}\\tan i. \\end{equation} Hence the observed $V_{\\rmn{LSR}}$ will be a linear function of either $x'$ or $y'$, if only one type of motion (rotation or expansion, respectively) is present in the ring \\citep{usc05}. Using equation (1), equation (4) can be written either in terms of the $x'$ or $y'$ coordinate as \\begin{equation} \\frac{[V_{\\rmn{LSR}}-v_s-(x'/a)v_{\\rmn{rot}}\\sin i]^2}{(v_{\\rmn{exp}}\\sin i)^2}+\\frac{x'^2}{a^2}=1, \\end{equation} \\begin{equation} \\frac{[V_{\\rmn{LSR}}-v_s-(y'/a)v_{\\rmn{exp}}\\tan i]^2}{(v_{\\rmn{rot}}\\sin i)^2}+\\frac{y'^2}{(a\\cos i)^2}=1. \\end{equation} Therefore, equations (5) and (6) indicate that the observed $V_{\\rmn{LSR}}$ has a quadratic form (ellipse) expressed in terms of $x'$ or $y'$, when both motions are present. In this model, we do not solve the radiative transfer through the ring, but we assume that the emission at a given $V_{\\rmn{LSR}}$ comes from the point of the ring having this line-of-sight velocity component. We then use this information to construct position-velocity diagrams. The orientation of the major axis of the ellipse in the position-velocity ($x'$-$V_{\\rmn{LSR}}$ or $y'$-$V_{\\rmn{LSR}}$) diagrams changes depending on the value of the inclination angle and on whether the sense of rotation is clockwise or counterclockwise as seen from the observer's point of view, as shown in Fig.~2. Accordingly, we are able to distinguish between a positive or negative value of the inclination angle and the sense of the rotation by doing a comparison between the position-velocity diagram delineated by the maser emission and the position-velocity diagrams expected for a rotating and expanding ring. It is important to note that similar position-velocity diagrams can be obtained considering contraction instead of expansion in the ring, but changing the sign of the inclination angle and the sense of rotation. This ambiguity can be solved by constraining the value of the inclination angle (positive or negative) with additional information (see Section 3.2). \\subsection[]{Least-squares fit} We carried out a least-squares fit of an ellipse to the observed emission, to estimate its spatial distribution on the sky. In order to do this, we considered the curve on the $xy$-plane given by equations (1)--(3) for a given set of values of the parameters $(x_0,y_0)$ (the centre of the ellipse on the plane of the sky), $a$ and $b$ (the semimajor and semiminor axes, respectively) and $\\theta$ (the angle between the $x$-axes of the two frames of reference). We then compute the minimum distances $d_j$ between the position $(x_j,y_j)$ of each of the masers and the ellipse (these distances are measured along straight lines that pass through the maser and intersect the ellipse at right angles to the curve). With these distances, we define the $\\chi^2$ as \\begin{equation} \\chi^2={1\\over {N-5}}\\sum_{j=1}^N \\left({d_j\\over \\sigma_j}\\right)^2\\,, \\label{chi2} \\end{equation} where $N$ is the number of data points (i.e. the factor $N-5$ is the number of degrees of freedom), and $\\sigma_j$ is the error associated with the position of the maser spots. We then find the set of parameters $(x_0,y_0,a,b,\\theta)$ (see above) which give the minimum of $\\chi^2$. As a first approximation, we began by setting the values of $\\sigma_j$ equal to the observational uncertainties $\\Delta_j$ for the measured positions of the masers, and then finding the ellipse that gives the minimum value of ${\\chi^2}$ (see equation \\ref{chi2}). The actual structure of the maser-emitting region could be a ring of finite width, for which a broad ellipse is a rough approximation of its projection on the plane of the sky. Therefore, we do not expect maser spots to exactly trace an ellipse. We can characterise the width of the ring, assuming that the fitted ellipse traces the mean projected angular distance of the maser emission to the central star, and the actual emission will be distributed around this ellipse, with a dispersion $\\Delta_e$. We treat $\\Delta_e$ as a source of error for the ellipse fit, additional to the measured error, so that ${\\sigma_j}^2={\\Delta_j}^2+{\\Delta_e}^2$. We then try different values of the width parameter ($\\Delta_e$), until we obtain a fit with a minimum ${\\chi^2}(\\Delta_e)=1$. We consider $2 \\Delta_e$ as the characteristic width of the maser ring. The $(x_0,y_0,a,b,\\theta)$ parameters obtained from the minimization of ${\\chi^2}$ give the best elliptical fit to the observed positions of the masers. With these spatial parameters, we carried out a kinematical fit, using the LSR velocities of the maser components in order to define a ${\\chi^2}_v$ of the form \\[ {\\chi^2}_v={1\\over {N-3}}\\sum_{j=1}^N {1\\over {{\\sigma_v}^2}} \\] \\begin{equation} \\times\\left(V_{\\rmn{LSR}},j-v_s-{{x_j}'\\over a}v_{\\rmn{rot}}\\sin i- {{y_j}'\\over a}v_{\\rmn{exp}}\\tan i\\right)^2\\,, \\end{equation} where $N-3$ is the number of degrees of freedom, and $({x'}_j,{y'}_j)$ are given by equations (2)--(3). Here $\\sigma_v$ is the uncertainty in the observed LSR velocity that we adopt as the spectral resolution of the observations. The minimization of ${\\chi^2}_v$ yielded the best values for the systemic ($v_s$), rotation ($v_{\\rmn{rot}}$), and expansion ($v_{\\rmn{exp}}$) velocities. \\begin{table*} \\centering \\begin{minipage}{112mm} \\caption{H$_2$O Masers in K3-35 (Region C)} \\label{symbols} \\begin{tabular}{@{}ccrcllccl} \\hline $V_{\\mathrm{LSR}}$ & \\multicolumn{3}{c}{Flux Density} & \\multicolumn{2}{c}{Position\\footnote{ Units of right ascension are hours, minutes, and seconds, and units of declination are degrees, arcminutes, and arcseconds. Data from \\citet{mir01} and \\citet{gom03}.}} & \\multicolumn{3}{c}{Position Uncertainty\\footnote{Relative position uncertainties (2$\\sigma$) between maser spots. The position of the 1.3~cm continuum emission peak is $\\alpha(\\rmn{J2000})=19^{\\rmn{h}}27^{\\rmn{m}}$44\\fs0233, $\\delta(\\rmn{J2000})=21\\degr30'$03\\farcs441. The relative position uncertainty between the continuum and the H$_2$O masers is 0\\farcs002. The accuracy of the absolute positions is 0\\farcs05.}} \\\\ (km~s$^{-1})$ & \\multicolumn{3}{c}{(mJy)} & $\\alpha$(J2000) & $\\delta$(J2000) & \\multicolumn{3}{c}{(arcsec)} \\\\ \\hline 24.6 & & 23 & & 19 27 44.0243 & 21 30 03.438 & & & 0.010 \\\\ 24.0 & & 61 & & 19 27 44.0242 & 21 30 03.428 & & & 0.004 \\\\ 23.3 & & 218 & & 19 27 44.02246 & 21 30 03.4460 & & & 0.0012 \\\\ 22.6 & & 1010 & & 19 27 44.022254 & 21 30 03.45146 & & & 0.00024 \\\\ 22.0 & & 1572 & & 19 27 44.022364 & 21 30 03.45335 & & & 0.00014 \\\\ 21.3 & & 945 & & 19 27 44.02247 & 21 30 03.4564 & & & 0.0003 \\\\ 20.7 & & 201 & & 19 27 44.02257 & 21 30 03.4625 & & & 0.0011 \\\\ \\hline \\end{tabular} \\end{minipage} \\end{table*} ", "conclusions": "From our model, we conclude that a ring is a likely explanation for the distribution and kinematics shown by the water masers located towards the centre of the PN K3-35. This ring may be arising from the innermost region of a disc or torus probably formed at the end of the AGB phase. Since masers trace regions with very stringent physical conditions, it is not possible to know whether the ring is tracing part of a toroidal or a disc-like structure. The kinematics of the ring in K3-35 suggests the presence of both rotating and expanding motions, as was also proposed in the young PN IRAS 17347-3139 \\citep{deg04}. The estimated expansion and rotation velocity values for K3-35 are similar (a few km~s$^{-1}$) to those obtained from water maser observations of IRAS 17347-3139. The calculated expansion velocity of the ring ($\\simeq$1.4~km~s$^{-1}$) in K3-35 is comparable to, but below, the expansion velocities of the tori inferred from interferometric CO observations in some PPNe and young PNe. For instance, the expansion velocities in M~1-92, M 2-9, M 2-56, and KjPn 8, have values in the range of $\\simeq$5--8~km~s$^{-1}$ \\citep{buj98,zwe97,cas02,for98}. It is noteworthy that the estimated expansion velocity in K3-35 obtained from water maser observations is quite similar to the slow expansion velocity ($\\simeq$0.8~km~s$^{-1}$) deduced from CO observations in the PPN Red Rectangle \\citep{buj05}. However, the estimated sizes of the tori detected in CO observations in those sources are about one order of magnitude more than the size of structure traced by the water masers in K3-35, suggesting that the structures observed in CO are probably related to the outermost equatorial region, while the water masers could be arising from the innermost one, such as a circumstellar disc, closer to the central star. \\begin{table} \\caption{Results of the rotating and expanding ring model fit to the H$_2$O masers towards the centre of K3-35.} \\label{symbols} \\begin{tabular}{@{}ccrcl} \\hline Parameter & \\multicolumn{4}{c}{Value$^a$} \\\\ \\hline $a$ & & 0\\farcs021 & $\\pm$ & 0\\farcs003 \\\\ $b$ & & 0\\farcs012 & $\\pm$ & 0\\farcs002 \\\\ $x_0\\,^*$ & & 0\\farcs001 & $\\pm$ & 0\\farcs001 \\\\ $y_0\\,^*$ & & 0\\farcs004 & $\\pm$ & 0\\farcs004 \\\\ PA & & 158$\\degr$ & $\\pm$ & 10$\\degr$ \\\\ $i$ & & 55$\\degr$ & $\\pm$ & 7$\\degr$ \\\\ $v_s$ & & 22.8 & $\\pm$ & 0.5~km~s$^{-1}$ \\\\ $v_{\\rmn{exp}}$ & & 1.4 & $\\pm$ & 0.9~km~s$^{-1}$ \\\\ $v_{\\rmn{rot}}$ & & 3.1 & $\\pm$ & 0.8~km~s$^{-1}$ \\\\ \\hline \\end{tabular} \\medskip $^a$ Note: Uncertainties are 2$\\sigma$.\\\\ $^*$ Coordinates of the centre of the ellipse relative to the position of the 1.3~cm continuum emission peak. \\end{table} In the case of K3-35, the water masers may be delineating an ellipse on the plane of the sky with a semimajor axis of $\\simeq$100 AU and a PA$\\simeq$158$\\pm$10\\degr, suggesting the presence of a disc or torus projected on the sky. The major axis of the ellipse is almost perpendicular (within the uncertainties) to the direction of the bipolar emission traced by the innermost region of the jet, with a PA$\\simeq$65\\degr observed by \\citet{mir01}, suggesting that the disc or torus could be physically related to the collimation of the bipolar outflow. Recently, \\citet{hug07} found that jets typically appear a few hundred years after the torus formation in PPNe and young PNe. This time sequence provides evidence that jets and tori are physically related. In the case of K3-35, \\citet{vel07} modelled its radio continuum emission as a precessing dense jet with an age of $\\sim$40~yrs. On the other hand, our estimate of the dynamical age of the ring traced by the maser emission is about $\\sim$350 years, using its radius ($\\sim$100~AU) and assuming a uniform expansion velocity of $\\sim$1.4~km~s$^{-1}$. These values would, in principle, suggest a lag in time similar to the value found by \\citet{hug07}. However, the dynamical age we derive is not reliable enough as an age estimate. Proper motion studies of the water maser emission (e.g. using e-MERLIN and VLBA) could provide better estimates. The kinematics of the ring in K3-35 suggests the presence of both expansion and rotation. A rough estimate for the central stellar mass can be obtained by assuming that the total energy (kinetic plus gravitational) of the masing gas is close to zero. In this case, $M\\simeq (R/2G)(v_{\\rmn{exp}}^2+v_{\\rmn{rot}}^2)$, where $R\\simeq100~(D/5~\\rmn{kpc})$~AU, is the radius of the ring. Hence, $M\\simeq[0.7~(D/5~\\rmn{kpc})\\pm 0.3]~\\rmn{M}_{\\sun}$, where $D$ is the source distance. The error in the mass only includes the errors in the fitted parameters. This estimate of the central mass is in agreement with the core mass required for a PNe, according to evolutionary models \\citep{kwo03}." }, "0808/0808.0896_arXiv.txt": { "abstract": "We present an observational study about the effects of the interactions in the kinematics, stellar population and abundances of the components of the galaxy pair AM\\,2306-721. Rotation curves for the main and companion galaxies were obtained, showing a deprojected velocity amplitude of 175 km s$^{-1}$ and 185 km s$^{-1}$, respectively. The interaction between the main and companion galaxies was modeled using numerical N-body/hydrodynamical simulations, with the result indicating that the current stage of the merger would be about 250 Myr after perigalacticum. The spatial variation in the distribution of the stellar population components in both galaxies was analysed by fitting combinations of stellar population models of different age groups. The central region of main galaxy is dominated by an old (5-10\\,Gyr) population, while significant contributions from a young (200\\,Myr) and intermediate (1\\,Gyr) components are found in the disk, being enhanced in the direction of the tidal features. The stellar population of the companion galaxy is overall much younger, being dominated by components with 1\\,Gyr or less, quite widely spread over the whole disk. Spatial profiles of the oxygen abundance were obtained from the a grid of photoionization models using the $R_{23}$ line ratio. The disk of the main galaxy shows a clear radial gradient, while the companion galaxy presents an oxygen abundance relatively homogeneous across the disk. The absence of an abundance gradient in the secondary galaxy is interpreted in terms of mixing by gas flows from the outer parts to the center of the galaxy due to the gravitational interaction with the more massive primary. ", "introduction": "It is widely accepted by some time that merging and interaction events play an important role in the formation and evolution of galaxies. Mergers change the mass function of galaxies, creating a progression from small galaxies to larger ones; the merging process can also change the morphology of the constituents, transforming gas-rich spirals in quiescent ellipticals. Interactions can also trigger a wide set of physical and morphological phenomena, such as tidal tails, bridges and shells, kinematically decoupled cores and star formation enhancements (see review of \\citealt{struck99}). Interacting galaxies also show enhanced star formation when compared with isolated galaxies. Such enhancement was initially proposed by \\citet{larson78} to explain the wider range of optical colors found in galaxies in pairs. Since then, numerous studies have confirmed these results, especially in the central regions, through measurements of optical emission lines \\citep{kennicutt84, kennicutt87, donzelli97, barton03, woods07}, infrared emission \\citep{joseph85, sekiguchi92, geller06} and radio emission \\citep{hummel81}. Recent studies have also shown that this enhancement is a function of the projected galaxy pair separation (e.g. \\citealt{barton00, lambas03, nikolic04}), being stronger in low-mass than in high-mass galaxies (e.g. \\citealt{woods07,ellison08}). In particular, \\citet{ellison08} found that the star formation rate as measured by the H$\\alpha$ equivalent width for galaxies in pairs selected from the Sloan Digital Sky Survey, is some 70\\% higher when compared to a control sample of galaxies with equal stellar mass distribution. There is a connection between the interaction strength and the morphological distortion in binary galaxies. According to \\citet{mihos96} models, the response of the gas to a close passage depends dramatically on the mass distribution of the galaxy, with the irregularities in the gas velocity field tracing the disturbances in the gravitational potential of the galaxy as observed, for example, in some galaxies in the Virgo cluster \\citep{rubin99}. Combined N-body/hydrodynamic simulations show that galaxy-galaxy mergers disturb the gas velocity field significantly, and hence lead to asymmetries and distortions in the rotation curves \\citep{kronberger06}. However, according to these authors, no severe distortions are observable about 1 Gyr after the first encounter. The gas motions created by the interaction can also significantly alter the chemical state of the galaxies \\citep{koeppen90,dalcanton07}, and modify the usually smooth radial metallicity gradient often found in isolated disk galaxies \\citep{henry99}. Recently, \\citet{kewley06} found that O/H abundance in the central region of nearby galaxy pairs is systematically lower that that of isolated objects. These authors suggest that the lower metallicity is a consequence of gas infall caused by the interaction, but very few observational studies have been published analysing in detail the influence of different levels of interactions in the metallicity distribution and enrichment properties of galaxies. Ferreiro \\& Pastoriza 2004 (hereafter FP04), in a study of the integrated photometry and star formation activity in a sample of interacting systems from the Arp-Madore catalogue \\citep{arp87}, found that the galaxies involved have bluer colours than those of isolated galaxies of the same morphological type, indicating an enhancement of star forming activity. This enhancement was also previously suggested by \\citet{donzelli97} to explain the slightly larger values of H$\\alpha\\, +\\, $\\ion{N}{ii} equivalent widths found in these systems, when compared with normal isolated spiral galaxies. From their sample, we have selected several systems to start a more comprehensive study of the effects of the interactions in the kinematics, stellar population and abundances of the galaxies in the so-called ``minor merger'' systems, defined here as physical pairs with mass ratio in the range of $0.04 < M_{secondary}/M_{primary} < 0.2 $. This paper presents the results for the system AM\\,2306-721. This pair is composed by a peculiar spiral with disturbed arms (hereafter, AM\\,2306A) in interaction with an irregular galaxy (hereafter, AM\\,2306B). Both galaxies contain very luminous \\ion{H}{ii} regions with H$\\alpha$ luminosity in the range of $8.30\\times\\,10^{39}\\,<\\,$L(H$\\alpha\\,)\\, <\\,1.27\\times\\,10^{42}$erg\\, s$^{-1}$ as estimated from H$\\alpha\\,$ images \\citep{ferreiro08}; and high star formation rate in the range of 0.07 to 10 $ M_{\\sun}$/yr \\citep{ferreiro08}. The present paper is organized as follows: in Section \\ref{datared}, we summarize the observations and data reduction. Section \\ref{vel} describes the gas kinematics of each galaxy and Section \\ref{numsim} present the numerical N-body/hydrodynamical simulations of the interaction. Section \\ref{sintese} presents the stellar population synthesis. The metallicity analysis is presented in Section \\ref{emission}, and the conclusions are summarized in Section \\ref{final}. ", "conclusions": "\\label{final} An observational study about the effects of the interactions in the kinematics, stellar population and abundances of the galaxy pair AM\\,2306-721 is performed. The data consist of long-slit spectra in the wavelength range of 3\\,350 to 7\\,130\\AA\\, obtained with the Gemini Multi-Object Spectrograph at Gemini South. The main findings are the following: \\begin{enumerate} \\item Rotation curves of the main and companion galaxies with an deprojected velocity amplitude of 175 km s$^{-1}$ and 185 km s$^{-1}$ respectively were obtained. An estimate of the dynamical mass was derived for each galaxy, using the deprojected velocity amplitude. For the main galaxy, its dynamical mass is $ 1.29 \\times 10^{11} M_{\\sun}$ within a radius of 18 kpc; and for the companion galaxy, the estimated dynamical mass is $ M(R)= 8.56 \\times 10^{10} M_{\\sun}$ within a radius of 10.7 kpc. \\item In the main galaxy, radial velocity deviations from the disk rotation of about 100 km\\,s$^{-1}$ were detected, which are probably due to the interaction with the companion galaxy. \\item In order to reconstruct the history of the AM\\,2306-721 system and to predict the evolution of the encounter, we modeled the interaction between AM\\,2306A and AM\\,2306B through numerical N-body/hydrodynamical simulations. The orbit that best reproduces the observational properties is found to be hyperbolic, with an eccentricity $e=1.15$ and perigalacticum of $q=5.25$ kpc; the current stage of the system would be about 250 Myr after perigalacticum. \\item The spatial variations of the stellar population components of the galaxies were analysed by fitting combinations of stellar population models of different ages (2.5 Myr, 200 Myr, 1 Gyr, 5 Gyr and 10 Gyr) and solar metallicity. The central regions of the main galaxy are dominated by the old (5-10\\,Gyr) population, with some significant contribution from a young 200\\,Myr and intermediate 1\\,Gyr component along the disk of the galaxy. On the other hand, the stellar population in the companion galaxy is overall much younger, being dominated by the 2.5\\,Myr, 200\\,Myr and 1\\,Gyr components, which are quite widely spread over the whole disk. \\item Oxygen abundance spatial profiles were obtained using a grid of photoionization models and the $R_{23}$=([\\ion{O}{ii}]$\\lambda\\,3727+$[\\ion{O}{iii}] $\\lambda\\,4959+$[\\ion{O}{iii}] $\\lambda\\,5007)/$H$\\beta$ line ratio. The disk of the main galaxy shows a clear radial oxygen abundance gradient, that can be fitted as a linear function $12+\\ohlog =8.75(\\pm0.06)-0.025(\\pm 0.007)R_{G}$; in the companion galaxy the oxygen abundance is relatively homogeneous across the galaxy disk, with the observed mean value of $12+\\ohlog =8.39$ \\item The absence of an abundance gradient in the secondary galaxy is interpreted as it having been destroyed by gas flows from the outer parts to the center of the galaxy. \\end{enumerate} \\thanks" }, "0808/0808.2294_arXiv.txt": { "abstract": "We show that repeated sound waves in the intracluster medium (ICM) can be excited by a single inflation episode of an opposite bubble pair. To reproduce this behavior in numerical simulations the bubbles should be inflated by jets, rather than being injected artificially. The multiple sound waves are excited by the motion of the bubble-ICM boundary that is caused by vortices inside the inflated bubbles and the backflow (`cocoon') of the ICM around the bubble. These sound waves form a structure that can account for the ripples observed in the Perseus cooling flow cluster. We inflate the bubbles using slow massive jets, with either a wide opening angle or that are precessing. The jets are slow in the sense that they are highly sub-relativistic, $v_j \\sim 0.01c-0.1c$, and they are massive in the sense that the pair of bubbles carry back to the ICM a large fraction of the cooling mass, i.e., $\\sim 1-50 M_\\odot \\yr^{-1}$. We use a two-dimensional axisymmetric (referred to as 2.5D) hydrodynamical numerical code (VH-1). ", "introduction": "\\label{sec:intro} Higher X-ray emissivity arcs, termed ripples, are observed in the intracluster medium (ICM) of the Perseus cluster (Fabian et al. 2003, 2006; Sanders \\& Fabian 2007), as well as in A2052 (Blanton et al. 2007). The ripples are thought to be sound waves that are excited by the inflation of bubbles near the cluster's center. Some of the ripples are observed on the outer edge of bubbles; they are compressed ICM that will turn into shock waves, and later to sound waves. It is possible that some of the other ripples are compressed by weak jets (Soker \\& Pizzolato 2005), rather than being sound waves detached from their parent bubble. Beside being interesting by their own merit, the sound waves can heat the ICM and partially offset the radiative cooling of the ICM (e.g., Br\\\"uggen et al. 2005; Fabian et al. 2005; Fujita \\& Suzuki 2005; Graham et al. 2008). \\par To obtain repeated sound waves previous numerical simulations assumed repeating bubble inflation episodes (e.g., Ruszkowski et al. 2004a,b; Sijacki \\& Springel 2006a, b). This had to be done because these simulations used \\emph{artificial bubbles}, i.e., numerically injected spherical bubbles at off-center locations, as is commonly done (e.g., Br\\\"uggen 2003; Br\\\"uggen \\& Kaiser 2001; Gardini 2007; Jones \\& De Young 2005; Pavlovski et al. 2008; Reynolds et al. 2005; Robinson et al. 2004; Ruszkowski et al. 2004a, 2004b, 2007, 2008; Scannapieco \\& Br\\\"uggen 2008). This of course is not the way bubbles are formed in clusters. Bubble are formed by jets [a different approach of injecting magnetic energy instead of a jet (Nakamura et al. 2008; Xu et al. 2008) requires further study]. When large, almost spherical bubbles (termed \\emph{fat bubbles}) are inflated by jets, some phenomena appear, that are not revealed when artificial bubbles are used (Sternberg \\& Soker 2008b). Among these are: (1) more stable bubbles, and (2) bubble-ICM boundary that is stochastically vibrates due to the vortices inside the bubble and the back flowing ICM. In previous works we found that in order for the jets to inflate fat bubbles they should have a large opening angle, or they should rapidly precess (Soker 2004, 2006; Sternberg et al. 2007; Sternberg \\& Soker 2008a). In addition, jets that are relatively slow, $v_j \\sim 10^4 \\km \\s^{-1} \\ll c$, and with a mass loss rate of the order of $2\\dot M_j \\sim1-50 M_\\odot \\yr^{-1}$ (for both jets), are more likely to inflate fat bubbles. The same results can be seen in the simulations of Omma et al. (2004), Alouani Bibi et al. (2007) and Binney et al. (2007), who use similar parameters to ours, but instead of conical jets with wide opening angles they use cylindrical jets with large initial radius. This is also shown by the simulation of Heinz et al. (2006), who manages to inflate a bubble (although not a bubble attached to the center) by launching a jet with a velocity of $3\\times 10^4 \\km \\s^{-1}$ and a mass loss rate of $35 M_\\odot \\yr^{-1}$ in one jet. In this paper we continue our study of jet-inflated bubbles, and show that such bubbles can excite several consecutive sound waves without the need to invoke periodic jet launching episodes. We do not deal with the propagation of sound waves and their properties. These seem to require more sophisticated treatment (Graham et al. 2008). We limit ourself to show that a proper treatment of bubble inflation can better explain the excitation of sound waves in the ICM. \\par ", "conclusions": "\\label{sec:discussion} We followed the response of the ICM to the inflation of bubbles by jets. The inflation of a bubble by a jet results in vortices inside the bubble and a backflow of the ICM around the bubble (Sternberg et al. 2007; Sternberg \\& Soker 2008a, b). Both processes cause the bubble-ICM to change shape with time and to become corrugated. In the case of a precessing jet (section \\ref{sec:precessing}), the changes in the jet axis cause a more prominent change in the bubble-ICM shape. This motion of the bubble-ICM boundary sends shocks and sound waves into the ICM. This effect cannot be reproduced by artificial bubbles, i.e., by numerically inserting spherical bubbles at off-center locations. \\par Our main result is that one episode of a bubble pair inflation can excite several sound waves along each radial direction. These form high-density arcs, the ripples. The front ripple is actually a weak shock. There is no need to introduce periodic (or semi-periodic) episodes of bubble inflation. Nonetheless, multiple episodes will make more ripples, and make the ripple structure more complicated. This is the case in Perseus, where the main central bubble pair was formed by jets that changed their direction and intensity (Dunn et al. 2006). We also observe a dense arc at the bubble's front that is apparent part of the time. The dense arc is the excitation of a new sound wave at the compression phase. Part of the time the region at the bubble's front is at the rarefaction phase, the wave trough, and no dense arc is observed there. \\par Our results can be incorporated into a broader scope. In a previous paper (Sternberg \\& Soker 2008b) we found that to follow the evolution of bubbles in the ICM one must inflate them by jets, rather than introduce them artificially. The evolution of bubbles and their influence on the ICM, e.g., sound waves, is crucial for the deposition of heat from the central active galactic nuclei to the ICM. Our works show that in order to understand the heating of the ICM one must properly inflate bubbles. \\par" }, "0808/0808.0772_arXiv.txt": { "abstract": "The effect of neutrino trapping on the longitudinal dielectric function at low densities has been investigated by using different relativistic mean field models. Parameter sets G2 of Furnstahl-Serot-Tang and Z271 of Horowitz-Piekarewicz, along with the adjusted parameter sets of both models, have been used in this study. The role of the isovector adjustment and the effect of the Coulomb interaction have been also studied. The effect of the isovector adjustment is found to be more significant in the Horowitz-Piekarewicz model, not only in the neutrinoless matter, but also in the matter with neutrino trapping. Although almost independent to the variation of the leptonic fraction, the instability region of matter with neutrino trapping is found to be larger. The presence of more protons and electrons compared to the neutrinoless case is the reason behind this finding. For parameter sets with soft equation of states at low density, the appearance of a large and negative $\\varepsilon_L (q,q_0=0)$ in some parts of the edge of the instability region in matter with neutrino trapping is understood as a consequence of the fact that the Coulomb interaction produced by electrons and protons interaction is larger than the repulsive isovector interaction created by the asymmetry between the proton and neutron numbers. ", "introduction": "\\label{sec_intro} At low densities both the relativistic and the non-relativistic mean field models predict a liquid-gas phase transition region for nuclear matter leading, for dense star matter, to a non-homogeneous phase commonly called pasta phase, which is formed by a competition between the long range Coulomb repulsion and the short range nuclear attraction~\\cite{Provi07}. This transition has substantial consequences on the properties of stellar matter and neutrino transport~\\cite{Duco07}. Considerable efforts to comprehend the uniform ground state stability of multi-component systems consisting of electrons, neutrinos, protons, and neutrons as a good approximation of this transition have been recently devoted, not only in the zero temperature approximation, but also for finite temperature~\\cite{Pethick,Douchin,Carr,Horowitz01,Provi1,Provi2,Provi3,Provi4,anto06,Muller:1995ji}. It is obvious that in order to understand the physics inside the non-homogeneous (unstable) regions like the mechanism of nuclear creation with slab-like or rod-like shape, we have to go beyond the mean field approximation. Attempts in this direction are discussed in Refs.~\\cite{Provi07,Napoli07,Watanabe04,Horowitz04,Sonoda}. Moreover, in the collapsing supernova core and at sub-nuclear densities, the transition of nuclear shape from sphere to other exotic shapes has significant effects to the neutrino mean free path. However, how these effects modify the neutrino mean free path is not fully understood yet~\\cite{Horowitz04,Sonoda}. Another motivation to study this transition comes from the fact that a neutron star is expected to have a solid inner crust of nonuniform neutron-rich matter above its liquid mantle~\\cite{Carr} and the mass of its crust depends sensitively on the density of its inner edge and on its equation of state (EOS)~\\cite{Douchin}. On the other hand, the critical density ($\\rho_c$), a density at which the uniform liquid becomes unstable to a small density fluctuation, can be used as a good approximation of the edge density of the crust~\\cite{Carr}. By generalizing the dynamical stability analysis of Ref.~\\cite{Horo1} in order to accommodate the various nonlinear terms in the relativistic mean field (RMF) model of Horowitz-Piekarewicz~\\cite{Horowitz01}, Carriere {\\it et al.}~\\cite{Carr} found a strong correlation between $\\rho_c$ in the neutron star and the density dependence of nuclear matter symmetry energy ($a_{\\rm sym}$). This leads to a suggestion that a measurement of the neutron radius in $^{208}{\\rm Pb}$ will provide useful information on the $\\rho_c$ \\cite{Carr,Horowitz01}. In our previous work~\\cite{anto06}, the critical densities of uniform matter with and without neutrino trapping have been calculated and analyzed by means of different RMF models. In this analysis it is shown that the interplay between the dominant contribution of the matter composition and the effective masses of mesons and nucleons leads to higher critical densities for matter with neutrino trapping. Furthermore, it was also found that the predicted critical density is insensitive to the number of trapped neutrinos as well as to the RMF model used. However, the discussion about the reason behind these findings was not quite robust. On the other hand, as we mentioned above, the neutrino transport is very crucial in the dynamics of the core-collapsing supernova due to the fact that the neutrinos carry most of the energy away and will lose also their energies by exciting collective nuclear and plasmon modes~\\cite{Provi2}. Similar situation can be also found in the neutron stars. Moreover, it was also shown in Ref.~\\cite{Provi2} that the behavior of electrons in matter depends strongly on the wave length or momentum of the external perturbation $q$. Note that this momentum is related to the energy transfer of the neutrinos that propagate in matter. The present paper reports on the extension of our previous investigation~\\cite{anto06} by calculating the longitudinal dielectric function of ERMF models and analyzing the relation between the obtained results and the isovector sector adjustment, the presence of the long-range Coulomb interaction, as well as the presence of electrons. The purpose of this work is to explain the reason behind the appearance of each point along the onset of the instability. To this end, we should emphasize here that we need to calculate the dielectric function in the edge of non-homogeneous regions because information from the critical density alone is insufficient. Furthermore, it is also important to emphasize that our definition of the instability is not the non-homogeneous area, but rather it is connected with the points where these non-homogeneities start to appear. As a consequence, the assumption of the uniform matter in the calculation is still valid. This paper is organized as follows. The RMF models and some constraints used in the present analysis are briefly discussed in Sec.~\\ref{sec_models}. In Sec.~\\ref{sec_LDF}, a discussion of the longitudinal dielectric function is given. In Sec.~\\ref{sec_results} we present the graphical results of the onset of instability along with the corresponding discussions. Finally, we give the conclusion in Sec.~\\ref{sec_conclu}. ", "conclusions": "\\label{sec_conclu} We have studied how the instability region starts to appear in low-density matter described by the Horowitz-Piekarewicz and Furnstahl-Serot-Tang models. To this end we have utilized the longitudinal dielectric function at $q_0=0$. The importance of the electron and Coulomb terms in matter with neutrino trapping has been investigated. It is found that the adjustment of the isovector terms has a more significant effect in the Horowitz-Piekarewicz model, i.e., producing a stronger repulsive isovector contribution which leads to a stronger correlation between its low density instability region and the $a_{\\rm sym}$ compared to the model of Furnstahl-Serot-Tang for matter without neutrino trapping. In the case of matter with neutrino trapping, the parameter sets with stiff EOS at low density lead to a large and positive $\\varepsilon_L (q,q_0=0)$. This demonstrates that, although the onsets of the instability of parameter sets with stiff and soft EOS at low densities are similar, the driving mechanisms are different. This fact might have an effect to the neutrino transport in matter. In both models the effect of the variation of the leptonic fraction is negligible, but the effect of the neutrino trapping on the onset of the instability region is significant. The presence of more protons and electrons in matter with neutrino trapping is the reason behind this phenomenon. The Coulomb term is found to be decisive in enlarging the stability of matter in this density region. The presence of the large and negative $\\varepsilon_L (q,q_0=0)$ in some parts of the instability region of matter with neutrino trapping originates from the fact that the isovector term is insufficient to cancel the attractive Coulomb interaction contributions generated by the presence of electrons (and protons)." }, "0808/0808.0919_arXiv.txt": { "abstract": "We performed cosmological, magneto-hydrodynamical simulations to follow the evolution of magnetic fields in galaxy clusters, exploring the possibility that the origin of the magnetic seed fields are galactic outflows during the star-burst phase of galactic evolution. To do this we coupled a semi-analytical model for magnetized galactic winds as suggested by \\citet{2006MNRAS.370..319B} to our cosmological simulation. We find that the strength and structure of magnetic fields observed in galaxy clusters are well reproduced for a wide range of model parameters for the magnetized, galactic winds and do only weakly depend on the exact magnetic structure within the assumed galactic outflows. Although the evolution of a primordial magnetic seed field shows no significant differences to that of galaxy clusters fields from previous studies, we find that the magnetic field pollution in the diffuse medium within filaments is below the level predicted by scenarios with pure primordial magnetic seed field. We therefore conclude that magnetized galactic outflows and their subsequent evolution within the intra-cluster medium can fully account for the observed magnetic fields in galaxy clusters. Our findings also suggest that measuring cosmological magnetic fields in low-density environments such as filaments is much more useful than observing cluster magnetic fields to infer their possible origin. ", "introduction": "Magnetic fields have been detected in galaxy clusters by radio observations, via the Faraday rotation signal of the magnetized cluster atmosphere towards polarized radio sources in or behind clusters \\citep{2002ARA&A..40..319C} and from diffuse synchrotron emission of the cluster atmosphere \\citep[see][for recent reviews]{2004IJMPD..13.1549G,2008SSRv..134...93F}. However, our understanding of their origin is still very limited. At present, models for the origin of seed fields can be classified in three main groups. In the first, a magnetic field is created in shocks through the ''Biermann battery'' effect \\citep{1997ApJ...480..481K,Ryu..1998,2001ApJ...562..233M}. A subsequent turbulent dynamo boosts it to the field strength observed in galaxy clusters. A second class of models invokes processes that took place in the early universe. In general, they predict that magnetic seed fields fill the entire volume of the universe; however the coherence length of the field crucially depends on the details of the models \\citep[see][for a review]{Grasso..PhysRep.2000}. Finally, galactic winds \\citep[e.g.][]{Volk&Atoyan..ApJ.2000} or AGN ejecta \\citep[e.g.][ and references therein]{1997ApJ...477..560E,Furlanetto&Loeb..ApJ2001} can produce magnetic fields and pollute the proto-cluster region. In such models, the magnetic field can also originate from an early population of dwarf, starburst galaxies \\citep{1999ApJ...511...56K} at relatively high redshift ($z \\approx 4 - 6$). In previous work, non radiative simulations of galaxy clusters within a cosmological environment which follow the evolution of a primordial magnetic seed field were performed using Smooth-Particle-Hydrodynamics (SPH) codes \\citep{1999A&A...348..351D,2002A&A...387..383D,2005JCAP...01..009D} as well as Adaptive Mesh Refinement (AMR) codes \\citep{2005ApJ...631L..21B,2008A&A...482L..13D,2008ApJS..174....1L}. Although these simulations are based on different numerical techniques they show good agreement in the predicted properties of the magnetic fields in galaxy clusters, when the evolution of an initial magnetic seed field is followed. This work has also demonstrated, that the properties of the final magnetic field in galaxy clusters do not depend on the detailed structure of the assumed initial magnetic field. The spatial distribution and the structure of the predicted magnetic field in galaxy clusters is primarily determined by the dynamics of the velocity field imprinted by cluster formation \\citep{1999A&A...348..351D,2002A&A...387..383D} and compares well with measurements of Faraday rotation. The creation of magnetic fields in shocks through the ''Biermann battery effect'' \\citep{1997ApJ...480..481K,Ryu..1998}, and subsequent turbulent dynamo action can be followed as well as a prediction can be made for magnetic field values from velocity fields inferred in cosmological simulations \\citep{2008Sci...320..909R}. Both methods predict magnetic field strengths in filaments with somewhat higher values \\citep[e.g. see ][]{2004PhRvD..70d3007S} than found in simulations that follow the evolution of a primordial magnetic seed field. Faraday rotation can be observed in several radio galaxies located at different radial distances with respect to the cluster center. Motivated by numerical simulations \\citep{2001A&A...378..777D}, the observed magnetic field is often modelled with a radially-declining field strength and a power law spectral structure. >From such observations, once can constrain the power law spectral index \\citep{2004A&A...424..429M,2006A&A...460..425G} or directly reconstruct the power spectrum of the magnetic field \\citep{2003A&A...412..373V,2005A&A...434...67V}. Given the sparse observational data available at the moment, a degeneracy exists between the central value of the magnetic field and its rate of radial decline \\citep[see for example][]{bonafede08,2008A&A...483..699G}, for which detailed predictions from simulations can be useful in breaking the degeneracy. The simulations must therefore examine different possible magnetic field origins in galaxy clusters in order to test the robustness of the inferred magnetic field properties. Recently the validity of models that produce a cluster magnetic field from galactic winds has been supported by a semi-analytic modelling of galactic winds \\citep{2006MNRAS.370..319B}. However, these models are unable to predict how the magnetic fields produced by the ejecta of the galaxies are compressed and amplified by the process of structure formation. Therefore, the structure of the final magnetic field in galaxy clusters cannot yet be predicted by these models. In this work we extend these studies by directly incorporating the galactic outflow model in magneto-hydrodynamical simulations of structure formation. The paper is structured as follows: In section \\ref{sim setup} we present the details of the cosmological setup, concentrating especially on the coupling of the semi-analytic model to the cosmological simulations. Details of the wind model and the seed magnetic field are presented in section \\ref{wind_model} and appendix \\ref{E2B}. The general results of our simulations are presented in section \\ref{results_all} and in section \\ref{results} we discuss our findings, particulary how galaxy clusters formed in our simulations. Finally, we present our conclusions in section \\ref{conclusions}. ", "conclusions": "\\label{conclusions} In this work we performed cosmological, magnetohydrodynamic simulations following the evolution of magnetic fields on large scale structures and in galaxy clusters. Coupling a semi-analytic model for magnetized galactic winds as suggested by \\citet{2006MNRAS.370..319B} with our cosmological simulation we explored the possibility that the magnetic fields in galaxy clusters originate from galactic outflows during star-burst phases, further processed by structure formation. We compared our results with the ones obtained by following a primordial magnetic seed field \\citep{2005JCAP...01..009D}. Performing several simulations, we explored the effect of various parameters of the adapted semi-analytic model relevant for the strength of the magnetic seed field from the galactic outflows. We also explored the effect of the magnetic field configuration assumed for the galactic outflows and of the seeding strategy. Our general findings are: \\begin{itemize} \\item The typical magnetic field strengths of several $\\mu$G in galaxy clusters as obtained from observations of Faraday rotation are well reproduced for a wide range of parameters of the galactic outflow model. \\item The general shape of the predicted Faraday rotation profile within clusters compares well with the sparse observational data available. Models that assume a field strength of $5\\mu$G within the galactic halo reproduce the observed Faraday rotation profiles better than models with a ten times stronger or a ten times weaker halo magnetic field. \\item The properties of the final magnetic field in galaxy clusters do not depend on the exact field configuration within the magnetic outflows. This confirms previous studies that the structure of the magnetic field in galaxy clusters is primarily driven by the velocity field induced by the structure formation process. \\item In massive galaxy clusters, the magnetic field amplification saturates around values of several $\\mu$G. The mass (or temperature) scale on which this happens depends on the strength of the magnetic seed field, and probably also on the resolution of the simulation. \\item In systems where saturation effects start to play a significant role, we observe only a small scatter in the magnetic scaling relations, and in the shape of the radial, magnetic profiles. \\item In clusters where saturation effects are negligible the strength and configuration of the magnetic field strongly depends on the dynamic state. Therefore, we observe a large scatter in the magnetic scaling relations and in the shape of the radial, magnetic profiles. \\item Within galaxy clusters, the structures predicted from synthetic Faraday rotation maps do not depend significantly on the magnetic seed field, and agree well with observed ones. \\item In low density environments imprints of the magnetic seed fields are still present and can be observed, in principle. In these environments the galactic outflows forming at later epochs contribute significantly to the magnetic field configuration. \\end{itemize} In summary, we confirm that galactic outflows and the subsequent action of structure formation can explain the properties of the magnetic fields observed in massive galaxy clusters. Their strength and shape is attributable to the velocity field induced by structure formation. We find that there are no measurable imprints of the magnetic seed fields left in the synthetic Faraday rotation maps of our simulated clusters. However, low density environments, like filaments, still contain significant information on the magnetic seed fields. Therefore, they may be used to discriminate proposed scenarios once information on magnetic fields within these regions becomes available with the next generation of radio instruments." }, "0808/0808.2139_arXiv.txt": { "abstract": "In the microquasar V4641 Sgr the spin of the black hole is thought to be misaligned with the binary orbital axis. The accretion disc aligns with the black hole spin by the Lense-Thirring effect near to the black hole and further out becomes aligned with the binary orbital axis. The inclination of the radio jets and the Fe\\ka\\ line profile have both been used to determine the inclination of the inner accretion disc but the measurements are inconsistent. Using a steady state analytical warped disc model for V4641 Sgr we find that the inner disc region is flat and aligned with the black hole up to about $900\\, R_{\\rm g}$. Thus if both the radio jet and fluorescent emission originates in the same inner region then the measurements of the inner disc inclination should be the same. ", "introduction": "Microquasars are binary systems in which material is accreted from a normal star on to a compact object. They differ from typical X-ray binaries by the strong presence of a persistent or episodic radio jet \\citep{MR99}. The compact object is usually associated with a neutron star or a stellar mass black hole. To date there are over fifteen microquasars for which the compact object has been dynamically confirmed to be a stellar mass black hole \\citep{O03}. However, only four of these systems have well resolved relativistic radio jets \\citep[XTE{\\thinspace J1550--564}, GRO{\\thinspace J1655--40}, GRS{\\thinspace 1915+105} and V4641 Sgr,][]{G03}. It is usually assumed that the inclination of the jet axis is perpendicular to that of the orbital plane of the binary. However, it has been shown by \\cite{Macc} and more recently by \\cite{MTP08} that the alignment time-scale in microquasars is usually a significant fraction of the lifetime of the system. If the black hole in such microquasars were formed with misaligned angular momentum, as expected from supernova-induced kicks, then it would be likely that the system would remain misaligned for most of its lifetime. Precise measurements of both the orbital plane and jet inclination are known for two systems. GRO{\\thinspace \\jb} is a microquasar thought to contain a rapidly rotating black hole \\citep{Z97,R08a} with a mass constrained to be larger than $6.0\\msun$ \\citep{OB97}. \\cite{H95} measured a jet inclination of $85$\\deg$\\pm{\\thinspace 2}$\\deg\\ to the line-of-sight and it is thus misaligned by at least $10$\\deg\\ to the orbital plane, with an inclination of $70.2$\\deg$\\pm{\\thinspace 1.9}$\\deg\\ \\citep{G01}. \\cite{MTP08} presented a detailed investigation of the alignment timescale of this system and found that it is consistent with the lifetime of the secondary star. V4641{\\thinspace Sgr} was discovered as an X-ray source independently with the Wide Field Cameras on \\bepo\\ on 1999 February 20 \\citep{intZ99} and with the Proportional Counter Array on the {\\it Rossi X-ray Timing Explorer} (\\rxte) on 1999 February 18 \\citep{MSM99}. Spectroscopic observations made between 1999 September 17 and 1999 October 16 led to a mass function $f(M) = 2.74\\pm0.12$\\msun\\ \\citep{O01}. From the lack of X-ray eclipses, combined with the large amplitude of the folded light curve, they deduced an orbital inclination angle in the range $60$\\deg$\\le i_{\\rm orbit} \\le 70$\\deg$.7$ and mass $8.73 \\le M_{\\rm BH} \\le 11.70$\\msun\\ for the compact object. Radio observations of \\vs\\ \\citep{H00} made during the 1999 September outburst found the jet expanding at apparent superluminal velocities with a proper motion ranging between $0.22\\arcsec-1\\arcsec$ per day. Based on the radio information, \\cite{O01} suggested that that the jet must be highly beamed and have an inclination along the line-of-sight of $i_{\\rm jet} \\le 10$\\deg. This differs significantly from from the inclination of the binary orbital axis. X-ray spectral analysis made on \\bepo\\ observations of \\vs\\ \\citep{intZ00} revealed the presence of a strong Fe\\ka\\ emission with an equivalent width between 0.3 and 1\\kev. They interpreted this as fluorescent emission using a photo-ionized medium. The presence of this Fe fluorescent emission was later confirmed by \\cite{R02} from a collection of \\rxte\\ data. They found an emission line at about $6.6$\\kev\\ with an equivalent width of about $360$\\ev. A more recent analysis of the source with \\rxte\\ data obtained during the outburst of 2003 August 5--17 \\citep{MB06} showed the presence of both a strong Fe\\ka\\ fluorescent emission line near 6.5\\kev\\ and a characteristic Compton hump at about $20$\\kev. If these features are attributed to the reprocessing of a hard X-ray (powerlaw) continuum by cold matter in an accretion disc \\citep{RN03,RF07} then the degree of broadening observed implies that the emitting region is very close to the black hole. The shape of the line profile can then give an indication of both the the radius of the emitting material from the black hole as well as the inclination of the inner accretion disc \\citep{F89,L91}. In this way \\cite{M02} obtained an estimate for the inclination of the innermost part of the accretion disc of $43$\\deg$\\pm{\\thinspace 15}$\\deg ($90\\%$ confidence). ", "conclusions": "We find that, in the accretion disc in V4641 Sgr, the region thought to be both the origin of the jets and the emission site of the Fe\\ka\\ line is flat and aligned with the central black hole. Thus we would expect the inclinations measured for the jets and with the Fe\\ka\\ line to be similar. Because there is a significant difference between the two measurements we conclude that one or both of them must be inaccurate or our model incorrect. It is important that this system be observed in more detail to resolve this in the near future." }, "0808/0808.1315.txt": { "abstract": " ", "introduction": "\\label{sect:intro} \\subsection{Scope of the report} \\label{sec:scope-document} The AMBER Task Force described in document [RD \\ref{rd:atfplan}] the objectives, its methodology and its plan to investigate the various AMBER issues which prevent the instrument to fulfill its initial specifications and therefore its original science program. The objectives of the February run was mainly to bring AMBER (see AMB-ATF-001 memo) into contractual specifications the accuracy of the absolute visibility, of the differential and of the closure phase through a fundamental analysis of the instrument status and limitations. \\subsection{Outline of the report} \\label{sec:outline-report} Before the run, a new implementation of the AMBER software by A.~Chelli and G.Duvert has been designed. This new and more accurate software using the same philosophy as the \\texttt{amdlib} v2.1\\footnote{\\texttt{amdlib} v2.1 is the last public release of the AMBER data reduction software.} is described in Sect.\\ \\ref{sec:dataproc}. The first days of the run were dedicated to the alignment of AMBER and characterization of its behavior. Many issues were tackled and the results are reported in Sect.\\ \\ref{sec:hardware}. Then we focused our attention on the main objective of our run: the performances limitations due to phase beating and to the lack of absolute calibration reported in Sect.\\ \\ref{sec:performances}. We end our report with the AMBER Task Force recommendations given in Sect.\\ \\ref{sec:recommendations}. \\emph{Note: we tried to keep the main part of the report as concise as possible and we moved the details in annexes.} \\subsection{Documents} \\subsubsection{Applicable documents} \\noindent \\begin{tabular}{p{6ex}p{0.45\\textwidth}p{0.37\\textwidth}} Code &Title &Number\\\\ \\hline \\ADlabel{ad:techspec} &AMBER Technical Specifications &VLT-SPE-ESO-15830-2074 issue 1.0 \\\\ &&dated 20.04.2000\\\\ \\hline \\end{tabular} \\subsubsection{Reference documents} \\noindent \\begin{tabular}{p{6ex}p{0.45\\textwidth}p{0.37\\textwidth}} Code &Title &Number\\\\ \\hline \\RDlabel{rd:atfplan} &AMBER Task Force: Objectives, Methodology, Plan &VLT-PLA-AMB-15830-7003 issue 1.0\\\\ &&dated 13/12/2008\\\\ \\RDlabel{rd:amb-atf-001} &ATF detailed plan for Feb'08 run &AMB-ATF-001 issue 1.5 dated 28/01/2008\\\\ \\RDlabel{rd:amb-igr-018} &AMBER data processing in the Image space &AMB-IGR-018 issue 1.0 dated 26/10/2000\\\\ \\RDlabel{rd:amb-det-007} &Report of the AMBER detector intervention from September 10 to 19, 2007 &AMB-DET-007 issue 1.0 dated 28/09/2007\\\\ \\RDlabel{rd:OPM-LLS} &OPM Warm Optics Design Report &VLT-TRE-AMB-15830-1001 issue 2.2\\\\ &&dated 28/05/2001\\\\ \\RDlabel{rd:OPM-TR} &OPM Warm Optics Test Report &VLT-TRE-AMB-15830-1010 issue 1.0\\\\ &&dated 10/10/2003\\\\ \\hline \\end{tabular} \\bigskip %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "" }, "0808/0808.2233_arXiv.txt": { "abstract": "We have studied superionization and X-ray line formation in the spectra of \\zpup\\ using our new stellar atmosphere code (XCMFGEN) that can be used to simultaneously analyze optical, UV, and X-ray observations. Here, we present results on the formation of the \\ion{O}{6}\\ $\\lambda\\lambda 1032, 1038$ doublet. Our simulations, supported by simple theoretical calculations, show that clumped wind models that assume void in the interclump space cannot reproduce the observed \\ion{O}{6} profiles. However, enough \\ion{O}{6}\\ can be produced if the voids are filled by a low density gas. The recombination of \\ion{O}{6} is very efficient in the dense material but in the tenuous interclump region an observable amount of \\ion{O}{6} can be maintained. We also find that different UV resonance lines are sensitive to different density regimes in \\zpup\\ : \\ion{C}{4} is almost exclusively formed within the densest regions, while the majority of \\ion{O}{6} resides between clumps. \\ion{N}{5} is an intermediate case, with contributions from both the tenuous gas and clumps. ", "introduction": "One of the surprising discoveries of the {\\it Copernicus} satellite was the strong P-Cygnii profiles of superions, such as \\ion{O}{6} and \\ion{N}{5}, in the FUV spectra of many O and B stars \\citep{sno76}. The only viable explanation for the presence of \\ion{O}{6} is Auger ionization by X-rays from \\ion{O}{4} \\citep{cas79} which is the dominant form of oxygen in many O-type stars. The X-ray emission, necessary for Auger ionization, was later detected by the first X-ray telescopes \\citep[e.g.,][]{har79, sew79}. The origin of the stellar X-ray emission was another enigma until the ``wind-shock'' mechanism \\citep{luc80} became the accepted explanation. Massive stars posses strong line-driven winds in which the material is accelerated by numerous C, N, O, and Fe transitions \\citep[see e.g.,][]{pau90, CAK}. It was known from the conception of the line-driven wind theory that such flows are unstable and prone to the formation of dense clumps and shocks \\citep[see e.g.,][]{owo91,luc80}. The large-scale flow energy is converted to heat in the shock fronts producing high temperature plasma. Numerical simulations confirm \\citep{fel97, owo88} that at least the soft X-ray emission of early-type stars can be explained by this mechanism. Evidence for density inhomogeneities (or clumped winds) is provided by variability studies of both WR stars \\cite[][and references therein]{lep99}, and O stars \\cite[e.g.,][]{eve98, lep08}. Further, density inhomogeneities allow the electron-scattering wings of emission lines to be reduced to the observed level while maintaining the strength of emission lines \\citep{hil91, ham98, hil99}. More recently, \\cite{cro02} and \\cite{hil03} found that they could not simultaneously fit the H$\\alpha$ and \\ion{P}{5} $\\lambda\\lambda$1120 profiles in normal O supergiants without assuming an inhomogeneous density distribution in the wind. Using a more statistical approach, \\cite{mas03} showed that the phosphorus ionization structure was consistent with expectations only if lower than conventional mass-loss rates were used in their analysis of \\ion{P}{5} $\\lambda\\lambda$1120. Additional observational evidence for wind clumping come from {\\it Chandra} and {\\it XMM-Newton} high-resolution X-ray spectra of O stars. These spectra revealed that X-ray lines suffer less absorption in the wind than predicted by ``smooth\" models \\citep[e.g.,][and references therein]{kra03b,coh05,wal07}. Superionization has received only limited attention since the work of \\citeauthor{mac93} \\citeyearpar{mac93, mac94}. The effect was introduced into modern stellar atmosphere codes \\citep[e.g., {\\it WM-basic},][]{pau01}, but we are unaware of any work that has revisited the question in light of the high-resolution X-ray observations, improved X-ray emission calculations and the new results on clumping. With improvements to CMFGEN, we are developing tools and techniques to move towards this goal. As part of this effort we discovered that the interclump medium is crucial to explain the \\ion{O}{6} doublet profile in $\\zeta$~Pup. In this paper we demonstrate the effect and discuss its implications. In \\S\\ref{sec:obs} we briefly describe our code, the observations we used, and our models. We present and discuss our results in \\S\\ref{sec:result}--\\S\\ref{sec:con}. ", "conclusions": "\\label{sec:con} In this letter we present our first results on superionization in clumped winds, and showcase the potential of the interclump medium to produce observable features. Clumped wind models that use the classical ``volume filling factor\" approach (clumps with voids in between) cannot reproduce the observed \\ion{O}{6} profile in $\\zeta$~Pup. The recombination of \\ion{O}{6} is too efficient and the necessary fractional abundance cannot be sustained in the clumps. However, a tenuous interclump medium can contribute enough \\ion{O}{6} to produce an observable \\ion{O}{6} profile. Only a small amount of mass is necessary in the interclump medium, so its overall effect on the derived mass-loss rates is negligible. Our result highlights the need for improved treatment of clumping in the winds of massive stars. It is impossible to achieve a simultaneous fit to all UV P-Cygnii profiles with a single component wind model for \\zpup. Our simulations suggest that in $\\zeta$~Pup, different UV resonance lines probe different density regimes. \\ion{C}{4} is formed almost exclusively in the dense material, while \\ion{O}{6} likely originates from the interclump medium. \\ion{N}{5} is an intermediate case with similar contributions from both components. In cooler O stars, when N$^{2+}$ become the dominant ionization stage, we might expect that \\ion{N}{5}\\ shows the same behavior as \\ion{O}{6}\\ in the hotter stars. Obviously other possible effects of the interclump medium, in both O and W-R stars, should be investigated." }, "0808/0808.2375_arXiv.txt": { "abstract": "We investigate general aspects of molecular line formation under conditions which are typical of prestellar cores. Focusing on simple linear molecules, we study formation of their rotational lines by radiative transfer simulations. We present a thermalization diagram to show the effects of collisions and radiation on the level excitation. We construct a detailed scheme (contribution chart) to illustrate the formation of emission line profiles. This chart can be used as an efficient tool to identify which parts of the cloud contribute to a specific line profile. We show how molecular line characteristics for uniform model clouds depend on hydrogen density, molecular column density, and kinetic temperature. The results are presented in a 2D plane to illustrate cooperative effects of the physical factors. We also use a core model with a non-uniform density distribution and chemical stratification to study the effects of cloud contraction and rotation on spectral line maps. We discuss the main issues that should be taken into account when dealing with interpretation and simulation of observed molecular lines. ", "introduction": "Star formation is a fundamental process in the universe. Dense, gravitationally bound, and genuinely starless cores, which we call here ``prestellar cores'', are the earliest visible precursors of forming stars. The interplay between gravitation and thermal/magnetic pressure as well as the conservation of angular momentum are all driving forces behind prestellar core evolution. While the thermal structure is determined by dust properties and molecular composition, magnetic support is sensitive to the ionization degree. The interaction and relative importance of these processes as well as the role of external forces are still not well understood \\citep[see, e.g. reviews by][]{Klessen:2004, Bergin:2007}. Thus, it is important to study the physical and chemical evolution of these cores in detail to reveal the underlying physics. The interiors of such prestellar cores are well-shielded from interstellar or stellar radiation, leading to low internal temperatures. Under such conditions, the main component of these cores, H$_2$, is not easily observable. Therefore, we have to rely on indirect methods to determine the physical structure of prestellar cores, e.g., on observations of thermal dust emission or emission lines from other molecules, like CO, CS and N$_2$H$^+$. Observations of spectral lines have an important advantage over continuum observations since they also carry information about kinematics of the gas. A disadvantage of molecular tracers can be that they are only present under certain conditions and can be frozen-out on dust surfaces. Numerous studies, including single-dish and interferometric observations along with their theoretical analysis, have been performed over the past years, significantly deepening our understanding of the physical and chemical processes in star-forming regions \\citep[see, e.g. reviews of][]{Myers:1999,Evans:1999,DiFrancesco:2007, Bergin:2007}. However, deriving physical properties from molecular lines is a difficult (inverse) problem because of the many factors which can affect the line formation. Even if we assume that the studied object is spherically symmetric and uniform, we have to specify at least 5 independent parameters describing the formation of molecular lines. These are density $n({\\rm H}_2)$, kinetic temperature $T_{\\rm kin}$, molecular column density $N({\\rm mol})$, radial velocity $V_r$, and micro-turbulent velocity $V_{\\rm turb}$. In more realistic models, additional parameters should also be considered, e.g., electron concentration and external radiation field, with spatial distributions of all the above parameters. An exact analytical treatment of this multi-parameter system is, in most cases, impossible. Optically thin line formation can be described analytically to some extent \\citep[see, e.g.,][]{Wilson:2000}, but it is very difficult to do the same for optically thick lines. Also, it is difficult to use analytic methods to solve the inverse problem, i.e., to restore physical distributions of relevant parameters from observed spectra because of the non-local and non-linear nature of the radiative transfer problem. This is why numerical line radiative transfer (LRT) models are commonly used as an exploratory tool. Several reliable numerical methods and numerical codes have been developed by various groups for this purpose \\citep[see reviews by][]{Peraiah:2004,Zadelhoff:2002}. There are also fast approximate numerical tools available for the molecular line radiative transfer analysis\\footnote{{\\it http://www.sron.rug.nl/\\~{}vdtak/radex/radex.php}} \\citep{vanderTak:2007}. However, input data for line modeling include not only physical conditions but also distributions of molecular abundances. All these data can be represented with some analytical prescription or extracted from chemical and dynamical models. As recently outlined by \\citet{tsamis}, there are two alternative approaches to the diagnostics of protostellar objects. The first approach is a systematic study of the influence of various factors on emergent spectra in order to facilitate the analysis of forthcoming observations. Such studies (based mainly on approximate LRT methods) have already started more than 30 years ago with the analysis of the excitation conditions for various molecules and effects of the density, temperature and velocity gradients on the emission line profiles \\citep[e.g.,][]{Lucas:1974,Goldreich:1974,Leung:1978,Stenholm:1980}. The alternative approach is a detailed study of an individual source. To infer its parameters, a number of models is constructed and the direct LRT problem is solved for each model. Varying the model parameters, it is possible to find the combination which provides the best agreement between observed and modeled spectra (maps) or their derived parameters, like, e.g., velocity centroids \\citep{walker1994} or line ratios \\citep{vanderTak:2007}. Both trial-and-error methods or any sophisticated numerical algorithm for isolating the ``best''\\ set of free parameters can be utilized \\citep{Keto:2004}. This second approach has already been successfully applied in a number of studies, aimed to extract detailed characteristics of pre-stellar and protostellar objects from observed spectra, in particular, their chemical and kinematical structure \\citep[e.g.,][]{Tafalla:2002,Tafalla:2004,Keto:2004, Evans:2005,Brinch:2007} as well as to test different star formation theories \\citep[e.g.][]{Pavlyuchenkov:2003, Offner:2007}. Following the first approach, efforts of various authors are mainly concentrated toward evolutionary stages later than the prestellar phase. Most cores investigated in detail are closer to the formation of a first hydrostatic core, when the collapse is well-developed and can be described by Shu or Larson-Penston solutions. This interest may be partly caused by the fact that first proofs of infall have been reported for Class\\,0 objects \\citep{walker1986,b335}, in which the collapse has already resulted in the formation of a deeply embedded source. \\cite{zhou1992} considered spectral differences between Shu and Larson-Penston solutions, assuming constant molecular abundances, and identified four line properties, supposedly unique to collapsing cores. \\citet{rawlings1992} combined a dynamical description, based on the Shu model, with a detailed chemical model and identified species with broad line wings, indicative of infall motions. \\citet{ry2001} coupled a similar chemical and dynamical model to a more realistic radiation transfer model, implemented as an approximate $\\Lambda$-iteration code, and studied the sensitivity of emergent central optically thin and optically thick spectra to various model parameters. The assumption of isothermality, used by \\citet{rawlings1992} and \\citet{ry2001}, was relaxed in \\citet{tsamis} in application to a ``B335-like'' Class~0 object. The sensitivity of line profiles in this model to the intensity of the ambient radiation field was studied in \\citet{redman2004}. The influence of various parameters on Class~0 central spectra, with a particular emphasis on turbulence and rotation, was also studied by \\citet{wtb2001} under the assumption of flat abundance distributions. The role of high angular resolution in studying such clouds was investigated by \\citet{choi}. More details about line modeling can be also found in review by \\citet{Evans:1999}. Large amounts of data have also been accumulated for starless cores, most of which are presumably ``prestellar'' \\citep[e.g.,][]{infall}, together with the theoretical modeling of molecular lines. E.g., \\citet{Tafalla:2004} investigated the effects of depletion on the line intensity. \\citet{Lee:2004} studied the evolution of line profiles during core contraction based on the models where the chemical evolution is calculated along with dynamical evolution of the cloud. \\citet{Pavlyuchenkov:2007} investigated the combined effects of temperature and depletion, and \\citet{DeVries:2005} investigated the possibility to use approximate analytical models to extract velocity gradients from the line profiles. In principle, all the mentioned studies of line formation in protostellar objects are also relevant for prestellar cores. However, despite the obvious observational and theoretical progress, the interpretation of the data is still far from being straightforward, as infall and rotation velocities in prestellar cores are both comparable or even smaller than the sound speed. Difficulties in restoring the information about structural, thermal, and kinematic properties are also caused by the lack of relevant methods for molecular line analysis. In this paper, we systematically study the formation of molecular lines in prestellar cores and present different tools which can be used to analyze the formation of lines in detail. Using the linear molecules CO and HCO$^+$ as examples, we model their emission by means of non-LTE LRT simulations. The role of hydrogen density, molecular column density, kinetic temperature, infall, and rotation is examined and illustrated in the paper. Such an analysis may form the basis of a more sophisticated interpretation of observational data and can be useful for those who are going to use LRT simulations in their studies or want to get a global view of the factors influencing observed line profiles. In Section~2 we describe the radiation transfer model used in the paper, introduce parameters to characterize line profiles, and provide some basic considerations on the line formation in prestellar cores. In Section~3 we show how molecular line characteristics depend on hydrogen density, molecular column density, and kinetic temperature for the parameter sample of uniform model clouds. In Section~4 we consider a non-uniform model cloud and study the effects of chemical stratification, contraction and rotation on spectral line maps. In Section~5 we discuss additional problems related to the LRT analysis. Section~6 summarizes the main conclusions of this paper. ", "conclusions": " \\begin{itemize} \\item Densities in starless cores fall into the range where level populations are neither radiatively, nor collisionally dominated. \\item Large column densities do not necessarily lead to the appearance of self-absorption dips. \\item When the column density is fixed, a specific line can be optically thick only in a range of densities, being optically thin at densities both below and above this range. \\item The density which is ``traced'' by some transition depends on external factors (UV field and CR ionization), which shape the molecular distribution. In particular, the ``traced'' density can be lower than the critical density. \\item Rotation and infall may produce very similar spectra and in general can only be distinguished by spectral mapping. \\end{itemize}" }, "0808/0808.2469_arXiv.txt": { "abstract": "We describe a new survey for unbound hypervelocity stars (HVSs), stars traveling with such extreme velocities that dynamical ejection from a massive black hole is their most likely origin. We investigate the possible contribution of unbound runaway stars, and show that the physical properties of binaries constrain low mass runaways to bound velocities. We measure radial velocities for HVS candidates with the colors of early A-type and late B-type stars. We report the discovery of 6 unbound HVSs with velocities and distances exceeding the conservative escape velocity estimate of Kenyon and collaborators. We additionally report 4 possibly unbound HVSs with velocities and distances exceeding the lower escape velocity estimate of Xue and collaborators. These discoveries increase the number of unbound HVSs by 60\\% - 100\\%. Other survey objects include 19 newly identified $z\\sim2.4$ quasars. One of the HVSs may be a horizontal branch star, consistent with the number of evolved HVSs predicted by Galactic center ejection models. Finding more evolved HVSs will one day allow a probe of the low-mass regime of HVSs and will constrain the mass function of stars in the Galactic center. ", "introduction": "Three-body interactions with a massive black hole (MBH) will inevitably unbind stars from the Galaxy \\citep{hills88, yu03}. In 2005 we reported the discovery of the first HVS: a 3 M$_{\\sun}$ main sequence star traveling with a Galactic rest frame velocity of at least $+700\\pm12$ km s$^{-1}$, more than twice the Milky Way's escape velocity at the star's distance of 110 kpc \\citep{brown05}. This star cannot be explained by normal stellar interactions: the maximum ejection velocity from binary disruption mechanisms \\citep{blaauw61, poveda67} is limited to $\\lesssim$300 km s$^{-1}$ for 3 M$_{\\sun}$ stars \\citep{leonard91, leonard93, tauris98, portegies00, davies02, gualandris05}. Although runaways may reach unbound velocities for very massive stars, like the hyper-runaway HD 271791 \\citep{heber08, przybilla08c}, runaway ejection velocities are constrained by the properties of binary stars. A massive and compact object is needed to accelerate low mass stars to unbound velocities. There is overwhelming evidence for a $\\sim4\\times10^6$ M$_{\\sun}$ MBH in the dense stellar environment of the Galactic center \\citep{schodel03, ghez08}. \\citet{hills88} coined the term HVS to describe a star ejected by the MBH. The observational signature of a HVS is its unbound velocity. Although not all unbound stars are necessarily HVSs -- fast-moving pulsars, for example, are explained by supernova kicks \\citep[e.g.][]{arzoumanian02} -- unbound low mass main sequence stars are most plausibly explained as HVSs. Here we introduce a new HVS survey using the MMT to target HVS candidates with masses down to $\\sim$2 M$_{\\sun}$. Discovering lower-mass HVSs should provide constraints on the stellar mass function of HVSs \\citep{brown06, kollmeier07}; the velocity distribution of low- versus high-mass HVSs may discriminate between a single MBH or binary MBH origin \\citep{sesana07b, kenyon08}. Our survey strategy targets stars fainter and redder than the original HVS survey. This strategy is successful: we report the discovery of 6 unbound HVSs and 4 possibly unbound HVSs. \\subsection{Recent Observations of HVSs} Observers have identified a remarkable number of HVSs in the past 3 years. Following the discovery of the first HVS \\citep{brown05}, \\citet{hirsch05} reported a helium-rich subluminous O star leaving the Galaxy with a rest-frame velocity of at least $+717$ km s$^{-1}$. \\citet{edelmann05} reported an 9 M$_{\\sun}$ main sequence B star with a Galactic rest frame velocity of at least $+548$ km s$^{-1}$, possibly ejected from the Large Magellanic Cloud. \\citet{brown06, brown06b, brown07a, brown07b} reported 7 B-type HVSs discovered in a targeted HVS survey, along with evidence for an equal number of bound HVSs ejected by the same mechanism. High-dispersion spectroscopy has shed new light on the nature of the HVSs. \\citet{przybilla08b} have recently shown that HVS7 is a chemically peculiar B main sequence star, with an abundance pattern unusual even for the class of peculiar B stars. The star HVS3 (HE 0437-5439), the unbound HVS very near the LMC on the sky, has received the most attention. HVS3 is a 9 M$_{\\odot}$ B star of half-solar abundance, a good match to the abundance of the LMC \\citep{bonanos08, przybilla08}. Stellar abundance may not be conclusive evidence of origin, however. A- and B-type stars exhibit 0.5 - 1 dex scatter in elemental abundances within a single cluster, due to gravitational settling and radiative levitation in the atmospheres of the stars \\citep{varenne99, monier05, fossati07, gebran08a, gebran08b}. An LMC origin requires that HVS3 was ejected from the galaxy at $\\sim$1000 km s$^{-1}$ \\citep{przybilla08}, a velocity that can possibly come from three-body interactions with an intermediate mass black hole in a massive star cluster \\citep{gualandris07, gvaramadze08}. \\citet{perets08b} shows that the ejection rate of 9 M$_{\\sun}$ stars, however, is four orders of magnitude too small for this explanation to be plausible. The alternative explanation is that HVS3 is a blue straggler, ejected by the Milky Way's MBH. Theorists argue that a single MBH or a binary MBH can eject a compact binary star as a HVS \\citep{lu07, perets08b}; the subsequent evolution of such a compact binary can result in mass-transfer and/or a merger that can possibly explain HVS3 \\citep{perets08b}. Proper motion measurements, underway now, will determine HVS3's origin. Other recent HVS work highlights the link between stellar rotation and the origin of HVSs. Main sequence B stars have fast mean $v\\sin{i}\\sim150$ km s$^{-1}$ \\citep[e.g.][]{abt02, huang06a}. Hot blue horizontal branch (BHB) stars have slow mean $v\\sin{i}<10$ km s$^{-1}$ (because they have just evolved off the red giant branch \\citep[e.g.][]{behr03, behr03b}). Interestingly, \\citet{hansen07} predicts that main sequence HVSs ejected by the Hills mechanism should be slow rotators, because stars in compact binaries have systematically lower $v\\sin{i}$ due to tidal synchronization. \\citet{lockmann08}, on the other hand, predict that HVSs should be fast rotators, at least for single stars spun up and ejected by a binary black hole. To date HVS1, HVS3, HVS7, and HVS8 have observed $v\\sin{i}$ of $\\sim$190, $55\\pm2$, $55\\pm2$, and $260\\pm70$ km s$^{-1}$, respectively \\citep{heber08b, przybilla08, przybilla08b, lopezmorales08}. As discussed by \\citet{perets07c}, we clearly require a larger sample of HVSs to measure the distribution of HVS rotations and discriminate HVS ejection models. In \\S 2 we describe our new HVS survey strategy and summarize our observations. In \\S 3 we discuss the Galactic escape velocity, our definition of a HVS, and the possible contribution of hyper-runaways to the population of unbound stars. In \\S 4 we present the new unbound HVSs. In \\S 5 we discuss a possible BHB star among the HVSs. We conclude in \\S 6. \\begin{figure}\t\t% \\plotone{f1.eps} \\caption{ \\label{fig:ugr} Color-color diagram illustrating the target selection for our new HVS survey ({\\it long dashed line}) and our old HVS survey ({\\it short dashed line}). The six new HVSs ({\\it solid stars}) and four possible HVSs ({\\it solid dots}) scatter around the \\citet{girardi04} stellar evolution tracks for 2 - 4 M$_{\\sun}$ main sequence stars ({\\it solid lines}). Average color uncertainties are indicated by the errorbar on the upper right. Previous HVS discoveries ({\\it open stars}) and the \\citet{xue08} BHB sample ({\\it x's}) are also marked.} \\end{figure} ", "conclusions": "We describe a new targeted HVS survey, a spectroscopic survey of faint stars $19+275$ km s$^{-1}$ and only 2 stars with $v_{rf}<-275$ km s$^{-1}$. Here we report the discovery of 6 new unbound HVSs in excess of the conservative escape velocity model of \\citet{kenyon08}, and 4 additional unbound HVSs in excess of the escape velocity model of \\citet{xue08}. One of the new HVSs may be an evolved BHB star. The \\citet{kenyon08} ejection models predict BHB HVSs are $\\sim$10 times less abundant than the main sequence HVSs in our survey, consistent with the existence of $1\\pm1$ BHB stars in our HVS sample. Of course, the exact number of BHB HVSs depends on the mass function of stars near the central MBH. BHB HVSs therefore have the potential to probe the low-mass regime of HVSs and constrain the mass function of stars in the Galactic center. HVSs are fascinating because their properties are tied to the nature and environment of the MBH that ejects them \\citep{levin06, baumgardt06, merritt06, ginsburg06, ginsburg07, demarque07, gualandris07, sesana06, sesana07, sesana07b, sesana07c, lu07, kollmeier07, hansen07, perets07, perets07c, perets08b, perets08a, sherwin08, svensson08, oleary08, lockmann08}, and their trajectories probe the dark matter halo through which they move \\citep{gnedin05, yu07, wu08, kenyon08}. The angular distribution of HVSs on the sky reveals significant anisotropy that may also be related to the Galactic potential \\citep{brown08d}. Our ultimate goal is to find a statistical sample of $\\sim$100 HVSs to measure the distribution of HVS properties and discriminate HVS ejection models. ~" }, "0808/0808.1140_arXiv.txt": { "abstract": "We report on the variability of 443 flat spectrum, compact radio sources monitored using the VLA for 3 days in 4 epochs at $\\sim 4$ month intervals at 5 GHz as part of the Micro-Arcsecond Scintillation-Induced Variability (MASIV) survey. Over half of these sources exhibited 2-10\\% rms variations on timescales over 2 days. We analyzed the variations by two independent methods, and find that the rms variability amplitudes of the sources correlate with the emission measure in the ionized Interstellar Medium along their respective lines of sight. We thus link the variations with interstellar scintillation of components of these sources, with some (unknown) fraction of the total flux density contained within a compact region of angular diameter in the range 10-50$\\mu$as. We also find that the variations decrease for high mean flux density sources and, most importantly, for high redshift sources. The decrease in variability is probably due either to an increase in the apparent diameter of the source, or a decrease in the flux density of the compact fraction beyond $z \\sim 2$. Here we present a statistical analysis of these results, and a future paper will the discuss the cosmological implications in detail. ", "introduction": "The discovery of centimeter-wavelength Intra-Day Variability (IDV) or ``flickering'' in Active Galactic Nuclei (AGN) by \\citet{hee84} initially raised concerns that some AGN possess brightness temperatures over six orders of magnitude above the $10^{12}\\,$K inverse Compton limit for incoherent synchrotron emission \\citep[e.g.][]{qui89}. However, considerable evidence has now accumulated to demonstrate that interstellar scintillation (ISS) in the turbulent, ionized interstellar medium (ISM) of our Galaxy is the principal mechanism responsible for the IDV observed in AGN, as was proposed by \\citet{HR87}. Two more recent observational techniques provide compelling evidence for the prevalence of ISS. Time delays of 1--8\\,min are observed in the arrival times of the flux density variations between telescopes on different continents for the three intra-hour variable sources B0405--385, B1257--326 and J1819$+$3845 \\citep{jau2000,big2006,dtdb2002}. The delay arises due to the finite time required for the stochastic fluctuations associated with the ISM to drift across the Earth. A second observational signature of ISS relates to the modulation of IDV variability timescales with a period of exactly one year. This arises because the Earth's orbital motion about the Sun contributes to the effective velocity with which the interstellar scattering material moves relative to an Earth-bound observer; the variations are slow as the Earth moves parallel to the material and fast as it moves anti-parallel to it. Annual cycles in IDV variability timescales are reported in at least seven sources \\citep[e.g.][]{den2003,ric2001,jau2001,big2003,jau2003}, including several whose long variability timescales preclude detection of time delays in the scintillation pattern over intercontinental distances. For many lines of sight through the ISM the slowest variations are expected in September if the motion of the turbulent material is comparable to the local standard of rest (LSR). The recognition of ISS as the dominant cause of IDV has not entirely alleviated the brightness temperatures problems posed by these sources. A source must be small to scintillate; in the weak scintillation case most frequently observed at frequencies near 5 GHz \\citep{wal98}, the source angular size must be comparable to or smaller than the angular size of the first Fresnel zone, $\\theta_{\\rm F}=\\sqrt{c/2 \\nu \\pi L}$. Here $L$ is the distance to the scattering region, which we will refer to as the screen even though in some cases it may be better described as a slab extending from the Earth out to distance $\\sim L$. $\\theta_{\\rm F}$ is typically tens of microarcseconds for screen distances of tens to hundreds of parsecs, which implies source components with angular sizes two to three orders of magnitude finer than the scales probed by VLBI. The long time-scale over which IDV has been observed in some sources suggests that such scintillating components can be relatively long-lived despite their small physical sizes. This paper reports on the results of a Micro-Arcsecond Scintillation-Induced Variability (MASIV) survey for IDV in AGN. This year-long survey conducted observations of between 500--700 AGN over each of four epochs of three or four days duration in 2002 and 2003 at 4.9\\,GHz with the VLA. The aim of the survey was to provide a catalogue of at least 100 AGN which vary on timescales of hours to days to provide the basis of detailed studies of the IDV AGN population drawn from a well-defined sample. A description of the observations in epochs 2, 3 and 4 is presented in \\S\\ref{obs} as a supplement to descriptions of the first epoch observations and MASIV source selection in \\citet{lov2003} (hereafter Paper 1). In \\S\\ref{VarClass} we describe how the time series of flux density for each source in each epoch was classified as variable or non-variable. In \\S\\ref{DsClass} we describe how we have quantified the amplitude and timescale for the variations from an analysis of the structure function combined from all 4 epochs and apply corrections for noise and other sources of flux density error. The basic hypothesis of the paper that the variations are predominantly due to interstellar scintillation is presented and examined in \\S\\ref{sec:ISS}, including the influence of parameters of the interstellar medium (\\S\\ref{VarB} emission measure and galactic latitude) and of the sources themselves (\\S\\ref{VarAlpha} mean flux density, spectral index). We now have redshifts for more than half of the sources and we present the dependence of the variability on redshift in \\S\\ref{VarZ} - a result which shows that the MASIV survey provides a new cosmological probe. In \\S\\ref{CompareVar} we consider whether the variability is intermittent over the four epochs. Our data are listed in Table \\ref{tab:BigTable} and our conclusions are presented in \\S\\ref{Conc}. ", "conclusions": "\\label{Conc} We have reported results from the four epochs of the MASIV survey. There were 710 sources with flat spectra ($\\alpha<-0.3$) near 5~GHz selected in weak and strong flux density groups surveyed for variability in four epochs over a year. These flat spectrum sources are predominantly quasars with compact emission probably from a core and jet, many with effective diameters small enough to show interstellar scintillation (ISS). In each epoch the flux density was measured using sub-arrays of the VLA every~2 hrs for about 12~hrs each day for 3-4~days. Sources were removed from the study if they showed evidence for changing correlated flux density due to confusion or resolution of their more extended structure, leaving 443 sources which were analyzed for variability in two ways. The first was a binary classification based on the raw modulation index (visual method) in which 43\\% of the sources were classified as variable in 2, 3 or 4 of the epochs. The second was a fit to the epoch-averaged structure function parameterized by $D(2\\mbox{d})$ and a timescale $\\tau_{\\rm char}$. In view of the uncertainties in the latter we classified sources as slow ($>3$~days), medium (0.5-3~days) or fast ($<0.5$~days) if $D(2\\mbox{d})$ exceeded $4\\times10^{-4}$. By this criterion 37\\% of the sources varied with more than 1.4\\% modulation index over 2~days, which is similar to the 43\\% variables by the visual classification. We found that $D(2\\mbox{d})$ and timescale varied both with coordinates in the Galaxy and also with source-based quantities. This confirms that the variations are dominated by ISS, which depends on both the strength of scattering and the distance to the scattering region and also on the fraction of flux density in its most compact component and its effective angular diameter. The following is a summary of our findings: \\begin{itemize} \\item The amplitude of 2-d variability increases with increasing emission measure estimated from H$\\alpha$ intensity for each line of sight. Emission measure is the column density for the square of the electron density which is expected to be strongly correlated with inhomogeneity in the ionized medium that causes ISS. This result provides observational evidence that ISS is the dominant cause of the variations. We find fast variations dominate for low emission measure, as expected since such regions will be seen out of the plane and closer to the Earth, and slow variations dominate for high emission measure which are typically seen at greater distances toward the Galactic plane especially for southern latitudes where the H$\\alpha$ intensity is high. \\item The amplitude of 2-d ISS variability varies significantly with Galactic latitude but differs substantially between positive and negative latitudes. The expected behaviour is complicated; greater path lengths at low latitudes, where the scattering should be stronger, cause the scattering to be slower which should reduce the rms over 3 d. However, the observed timescales show that there are more sources with fast variations at high latitudes and more sources with slow variations at low latitudes in both hemispheres, in clear support of ISS as the dominant cause. \\item The ISS modulation index tends to decrease with increasing mean flux density, as expected if the compact emission is limited by synchrotron self absorption or inverse Compton losses to have a maximum brightness temperature. In that case the expected angular diameter $\\propto \\sqrt{\\bar{S}}$ which will quench the ISS of the stronger sources. \\item There is little change in the ISS amplitude with spectral index for our sample with $\\alpha > -0.3$. \\item There is evidence that the ISS can be intermittent on times of 3-6 months for some sources, but this is hard to quantify from the 3~day observing sequences, when the time scale of the variations is of the same order. \\item We model $D(2\\mbox{d})$ as a function of compact source component fractional flux density and angular diameter, from which we find compact diameter to lie in the range 0.005 -- 0.15 milli arcseconds and brightness temperatures in the range $10^{12} - 10^{14}$K. \\item The most far-reaching result reported here is the discovery of a decrease in both the fraction of sources that scintillate and in their scintillation amplitude beyond redshifts around 2. We conclude that there is an increase in the typical angular diameter of the most compact radio-emitting regions of the quasars beyond reshift 2. The possible interpretations of this exciting result will be presented in a companion paper \\citep{macq08}. \\item A further surprise (at least to us) was the apparent absence of the very rapid variables (IHV). J1819+3845 fell in our sample, but it was the only source to show remarkable large amplitude variability. J0929+5013 showed rapid variability in the January 2002 epoch \\citep{lov2003} but, although monitored closely, revealed only slower, many-hour variability in the three later epochs. We had expected to find more of these rapid variables especially given that two of the three known, J1819+3845 and PKS1257-326 were discovered serendipitously. This strongly suggests that the IHV sources lie behind discrete local interstellar clouds which cover a small fraction of the sky. \\end{itemize}" }, "0808/0808.2432_arXiv.txt": { "abstract": "Spherical models of collisionless but quasi-relaxed stellar systems have long been studied as a natural framework for the description of globular clusters. Here we consider the construction of self-consistent models under the same physical conditions, but including explicitly the ingredients that lead to departures from spherical symmetry. In particular, we focus on the effects of the tidal field associated with the hosting galaxy. We then take a stellar system on a circular orbit inside a galaxy represented as a ``frozen\" external field. The equilibrium distribution function is obtained from the one describing the spherical case by replacing the energy integral with the relevant Jacobi integral in the presence of the external tidal field. Then the construction of the model requires the investigation of a singular perturbation problem for an elliptic partial differential equation with a free boundary, for which we provide a method of solution to any desired order, with explicit solutions to two orders. We outline the relevant parameter space, thus opening the way to a systematic study of the properties of a two-parameter family of physically justified non-spherical models of quasi-relaxed stellar systems. The general method developed here can also be used to construct models for which the non-spherical shape is due to internal rotation. Eventually, the models will be a useful tool to investigate whether the shapes of globular clusters are primarily determined by internal rotation, by external tides, or by pressure anisotropy. ", "introduction": "Large stellar systems can be studied as collisionless systems, by means of a one-star distribution function obeying the combined set of the collisionless Boltzmann equation and the Poisson equation, under the action of the self-consistent mean potential. For elliptical galaxies the relevant two-star relaxation times do actually exceed their age; an imprint of partial relaxation may be left at the time of their formation \\citetext{if we refer to a picture of formation via incomplete violent relaxation; \\citealp{Lyn67}, \\citealp{Alb82}}, but otherwise they should be thought of as truly collisionless systems, generally characterized by an anisotropic pressure tensor. In turn, for globular clusters the relevant relaxation times are typically shorter than their age, so that we may argue that for many of them the two-star relaxation processes have had enough time to act and to bring them close to a thermodynamically relaxed state, with their distribution function close to a Maxwellian. This line of arguments has led to the development of well-known dynamical models for globular clusters \\citep{Kin65,Kin66}. King models are based on a quasi-Maxwellian isotropic distribution function $f_K(E)$ in which a truncation prescription, continuous in phase space, is set heuristically to incorporate the presence of external tides; but otherwise, they are fully self-consistent (i.e., no external fields are actually considered) and perfectly spherical. Empirically, the simplification of spherical symmetry is encouraged by the fact that in general globular clusters have round appearance. Indeed, as a zero-th order description, these models have had remarkable success in applications to observed globular clusters \\citep[e.g., see][and references therein]{Spi87, DjoMey94,McLMar05}. In recent years, great progress has been made in the acquisition of detailed quantitative information about the structure of these stellar systems, especially in relation to the measurement of the proper motions of thousands of individual stars \\citep[see][]{Lee00,McL06}, with the possibility of getting a direct 5-dimensional view of their phase space. Such progress calls for renewed efforts on the side of modeling. More general models would allow us to address the issue of the origin of the observed departures from spherical symmetry. In fact, it remains to be established which physical ingredient among rotation, pressure anisotropy, and tides is the primary cause of the flattening of globular clusters \\citep[e.g., see][and references therein]{Kin61, FreFal82,Gey83,WhiSha87,DavPru90,HanRyd94,Ryd96,Goo97,Ber08} As in the case of the study of elliptical galaxies \\citep[e.g., see] [and references therein]{BerSti93}, different approaches can be taken to the construction of models. Broadly speaking, two complementary paths can be followed. In the first, ``descriptive\" approach, under suitable geometrical (on the intrinsic shape) and dynamical (e.g., on the absence or presence of dark matter) hypotheses, the available data for an individual stellar system are imposed as constraints to derive the internal orbital structure (distribution function) most likely to correspond to the observations. This approach is often carried out in terms of codes that generalize a method introduced by \\citet{Sch79}; for an application to the globular cluster $\\omega$ Cen, see \\citet{Ven06}. In the second, ``predictive\" approach, one proposes a formation/evolution scenario in order to identify a physically justified distribution function for a wide class of objects, and then proceeds to investigate, by comparison with observations of several individual objects, whether the data support the general physical picture that has been proposed. Indeed, King models belong to this latter approach. The purpose of this paper is to extend the description of quasi-relaxed stellar systems, so far basically limited to the spherical King models, to the non-spherical case. There are at least three different ways of extending spherical isotropic models of quasi-relaxed stellar systems (such as King models), by modifying the distribution function so as to include: (i) the explicit presence of a non-spherical tidal field; (ii) the presence of internal rotation; (iii) the presence of some pressure anisotropy. As noted earlier in this Introduction, these correspond to the physical ingredients that, separately, may be thought to be at the origin of the observed non-spherical shapes. We will thus focus on the construction of physically justified models, as an extension of the King models, in the presence of external tides and, briefly, on the extension of the models to the presence of rigid internal rotation. A first-order analysis of the triaxial tidal problem addressed in this paper was carried out by \\citet{HegRam95} and the effect of a ``frozen'' tidal field on (initially) spherical King models was studied by \\citet{Wei93} using N-body simulations. For the axisymmetric problem associated with the presence of internal rotation, a first-order analysis of a simply truncated Maxwellian distribution perturbed by a rigid rotation was given by \\citet{KorAna71}; different models where differential rotation is considered were proposed by \\citet{PreTom70} and by \\citet{Wil75}, also in view of extensions to the presence of pressure anisotropy, which goes beyond the scope of this paper. Models that represent a direct generalization of the King family to the case with differential rotation have also been examined, with particular attention to their thermodynamic properties \\citep{LagLon96,LonLag96}. In principle, the method of solution that we will present below can deal with the extension of other spherical isotropic models with finite size that are not of the King form \\citetext{e.g., see \\citealp{WooDic62,Dav77}; see also the interesting suggestion by \\citealp{Mad96}}. The study of self-consistent collisionless equilibrium models has a long tradition not only in stellar dynamics, but also in plasma physics \\citep[e.g., see][]{Har62,AttPeg99}. We note that in both research areas a study in the presence of external fields, especially when the external field is bound to break the natural symmetry associated with the one-component problem, is only rarely considered. The paper is organized as follows. Section 2 introduces the reference physical model, in which a globular cluster is imagined to move on a circular orbit inside a host galaxy treated as a frozen background field; the modified distribution function for such a cluster is then identified and the relevant parameter space defined. In Sect.~3 we set the mathematical problem associated with the construction of the related self-consistent models. For models generated by the spherical $f_K(E)$, Sect.~4 and Appendix A give the complete solution in terms of matched asymptotic expansions. Alternative methods of solutions are briefly discussed in Sect.~5. The concluding Section 6 gives a summary of the paper, with a short discussion of the results obtained. In Appendix B we show how the method developed in this paper can be applied to construct quasi-relaxed models flattened by rotation in the absence of external tides. In Appendix C we show how the method can be applied to other isotropic truncated models, different from King models. Technically, the mathematical problem of a singular perturbation with a free boundary that is faced here is very similar to the problem noted in the theory of rotating stars, starting with Milne \\citep[see][]{Tas78,Mil23,Cha33,Kro42,ChaLeb62,MonRox65}. The problem was initially dealt with inadequate tools; a satisfactory solution of the singular perturbation problem was obtained only later, by \\citet{Smi75,Smi76}. ", "conclusions": "Spherical King models are physically justified models of quasi-relaxed stellar systems with a truncation radius argued to ``summarize\" the action of an external tidal field. Such simple models have had great success in representing the structure and dynamics of globular clusters, even though the presence of the tidal field is actually ignored. Motivated by these considerations and by the recent major progress in the observations of globular clusters, in this paper we have developed a systematic procedure to construct self-consistent non-spherical models of quasi-relaxed stellar systems, with special attention to models for which the non-spherical shape is due to the presence of external tides. The procedure developed in this paper starts from a distribution function identified by replacing, in a reference spherical model, the single star energy with the relevant Jacobi integral, thus guaranteeing that the collisionless Boltzmann equation is satisfied. Then the models are constructed by solving the Poisson equation, an elliptic partial differential equation with free boundary. The procedure is very general and can lead to the construction of several families of non-spherical equilibrium models. In particular, we have obtained the following results: \\begin{itemize} \\item We have constructed models of quasi-relaxed triaxial stellar systems in which the shape is due to the presence of external tides; these models reduce to the standard spherical King models when the tidal field is absent. \\item For these models we have outlined the general properties of the relevant parameter space; in a separate paper (Varri \\& Bertin, in preparation) we will provide a thorough description of this two-parameter family of models, also in terms of projected quantities, as appropriate for comparisons with the observations. \\item We have given a full, explicit solution to two orders in the tidal strength parameter, based on the method of matched asymptotic expansions; by comparison with studies of analogous problems in the theory of rotating polytropic stars, this method appears to be most satisfactory. \\item We have also discussed two alternative methods of solution, one of which is based on iteration seeded by the spherical solution; together with the use of dedicated $N$-body simulations, the ability to solve such a complex mathematical problem in different ways will allow us to test the quality of the solutions in great detail. \\item By suitable change of notation and physical re-interpretation, the procedure developed in this paper can be applied to the construction of non-spherical quasi-relaxed stellar systems flattened by rotation (see Appendix B). \\item The same procedure can also be applied to extend to the triaxial case other isotropic truncated models (such as low-$n$ polytropes), that is models that do not reduce to King models in the absence of external tides (see Appendix C). \\end{itemize} We hope that this contribution, in addition to extending the class of self-consistent models of interest in stellar dynamics, will be the basis for the development of simple quantitative tools to investigate whether the observed shape of globular clusters is primarily determined by internal rotation, by external tides, or by pressure anisotropy." }, "0808/0808.2574_arXiv.txt": { "abstract": "To reconstruct dark energy models the redshift $z_{eq}$, marking the end of radiation era and the beginning of matter-dominated era, can play a role as important as $z_{t}$, the redshift at which deceleration parameter experiences a signature flip. To implement the idea we propose a variable equation of state for matter that can bring a smooth transition from radiation to matter-dominated era in a single model. A popular $\\Lambda \\propto \\rho$ dark energy model is chosen for demonstration but found to be unacceptable. An alternative $\\Lambda \\propto \\rho a^{3}$ model is proposed and found to be more close to observation. ", "introduction": "The current trend in modern cosmology is mainly focussed on issues of possible explanations of the observed late time acceleration in the expansion of our Universe\\citep{riess,perl}. The presence of a dark energy of unknown origin which dominates over matter in our Universe is now almost accepted. Unlike ordinary matter the dark energy has a repulsive effect and its dominance accounts for the observed acceleration. Several candidates of dark energy are found in literature. Among them the most elementary is the cosmological constant $\\Lambda$\\citep{padd,peebles}. Variable $\\Lambda$ models\\citep{car,wet,arbab,paddy,vis,shap,dutta} were introduced later in an effort to overcome the fine tuning problem. Among other candidates most popular are different kinds of scalar fields like quintessence\\citep{caldwell1}, K-essence\\citep{chiba}and phantom fields\\citep{caldwell2} to name only a few. The Chaplygin gas\\citep{kamen} has an equation of state that can produce negative pressure in later phase of evolution to account for the accelerated expansion of the Universe. Some geometrical origin of the late-time acceleration are also found in modified gravity models\\citep{nojiri}. All these models have their own merits and demerits. The exact nature of dark energy is therefore yet to be ascertained. For a detailed review on different dark energy candidates one may see the work of \\citet{sahni} or \\citet{copeland}. Most of these models are usually studied only in the matter dominated phase assuming the presence of some kind of dark energy to explain the transition from deceleration to acceleration. This is due to the fact that radiation era was very short lived and dominated the Universe in its early phase of evolution whereas acceleration is a more recent phenomena which has occurred very late in the matter dominated epoch. The viability of such models are usually checked by comparing the variation of deceleration parameter $q$ with the red-shift $z$ obtained theoretically with that from observations. A good estimation of $z=z_t$, the redshift marking the transition from deceleration to acceleration, is available from observation. But observational data are limited over a small range of $z$ values around $z_t$. Hence many models that fit observational data are found to behave quite differently outside this range. To choose among these models we need at least another event that happens to be outside this range of $z$, preferably in the distant past. We point out that a very good estimate of $z=z_{eq}$, the redshift when matter density was equal to radiation density, is at our disposal. Thus instead of a single point we now have two well separated points through which the $q$ vs. $z$ curve for any viable cosmological model should pass. Such a model would be more reliable as its viability could be ensured over a large range of $z$ values. But then to estimate $z_{eq}$ we must have both radiation and matter in a single model. \\par In the next section we propose a variable equation of state for matter that mimics radiation in the past and pressureless dust in the later course of evolution. In section 3 we examine the $\\Lambda \\propto \\rho$ model with the proposed equation of state for matter. In section 4 an alternative variable $\\Lambda $ model is proposed. ", "conclusions": "We have shown in this present article that the viability of a dark energy model can be cheked in a better way by demanding that $z_{eq}$ should be $\\sim 10^{3}$. Introducing a variable equation of state for matter we have investigated the nature of expansion of the Universe in two variable dark energy models. The variable equation of state for matter ensures that the early epoch of the Universe is dominated by radiation. With the expansion of the Universe the pressure falls as $a^{-m}$ $(m>0)$ to produce an almost pressure-free late-Universe. The equation of state for dark energy is essentially chosen to be $\\alpha_{DE}=p_{DE}/\\rho_{DE}=-1$ while the $\\rho_{DE}$ is in general a function of the scale factor $a$. This in other way amounts to the same as the presence of a dynamical $\\Lambda$ in the field equations. It makes the conservation equation notably different. One has to allow interaction among the matter part and dark energy. The functional form of $\\Lambda$ is then chosen to produce late-time acceleration in the expansion of the Universe. A $\\Lambda \\propto \\rho$ model fails to meet observational constraints. A choice of $\\Lambda \\propto \\rho a^{3}$ is found to be a better alternative. Thus we conclude that meeting the additional $z_{eq}\\sim 10^{3}$ requirement may be an essential feature of reconstructing dark energy models." }, "0808/0808.2742_arXiv.txt": { "abstract": "We report the discovery of the largest giant radio galaxy, J1420-0545: a FR type II radio source with an angular size of 17.4' identified with an optical galaxy at $z$=0.3067. Thus, the projected linear size of the radio structure is 4.69 Mpc (if we assume that $H_{0}$=71 km\\,s$^{-1}$Mpc$^{-1}$, $\\Omega_{\\rm m}$=0.27, and $\\Omega_{\\Lambda}$=0.73). This makes it larger than 3C236, which is the largest double radio source known to date. New radio observations with the 100 m Effelsberg telescope and the Giant Metrewave Radio Telescope, as well as optical identification with a host galaxy and its optical spectroscopy with the William Herschel Telescope are reported. The spectrum of J1420$-$0545 is typical of elliptical galaxies in which continuum emission with the characteristic 4000\\,\\AA\\hspace{1mm}discontinuity and the $H$ and $K$ absorption lines are dominated by evolved stars. The dynamical age of the source, its jets' power, the energy density, and the equipartition magnetic field are calculated and compared with the corresponding parameters of other giant and normal-sized radio galaxies from a comparison sample. The source is characterized by the exceptionally low density of the surrounding IGM and an unexpectedly high expansion speed of the source along the jet axis. All of these may suggest a large inhomogeneity of the IGM. ", "introduction": "The existence of bright hot spots at the edges of most of FR type II radio sources indicates that the jets ejected from a \"central engine\" in the AGN encounter resistance to their propagation. These supersonically advancing hotspots are likely to be confined by the ram pressure of the external environment, which on the largest scales is the intergalactic medium (IGM). The shocked jet material and shocked IGM form a cocoon surrounding the pair of jets. In a number of FR type II radio sources, such a cocoon is observed as a low-brightness bridge between the radio lobes \\citep{leahy}. While the lengthening of the cocoon is governed by the balance between the jet's thrust and the ram pressure of the IGM, the width of the cocoon is determined by its mean internal pressure, which is on the order of the kinetic energy delivered by the jets divided by the cocoon's volume. As both of these factors are functions of time, it is possible that the lateral expansion of the cocoon could be faster than the rate at which the jets delivered energy. Thus, the cocoon's pressure would decrease until it reached an equilibrium with the IGM pressure. However, in a number of papers (e.g. \\citealt{begel}; \\citealt{daly}; \\citealt{wdw}) there are arguments that the cocoons in many observed radio sources are still overpressured with respect to the IGM. Only the very tenuous material in the bridges of the oldest and largest sources is expected to have attained such an equilibrium state, in which the radiating particles and the magnetic field are balanced, and the pressure of the relativistic plasma equals the pressure of the gaseous environment (cf. \\citealt{nath}). Consequently, a determination of the physical conditions in these diffuse bridges of large-size double radio sources (as they are far away from the ram pressure-confined heads of the source) gives us an independent tool with which to probe the pressure of the IGM. The above approach was applied by us \\citep{mkj} to a sample of giant radio galaxy (GRG) candidates located in the southern sky hemisphere with the aim of studying properties of the cosmological evolution of the IGM. The radio galaxy discussed in this paper is a member of that sample. The source J1420$-$0545 was discerned in the VLA 1.4 GHz surveys FIRST \\citep{becker} and NVSS \\citep{condon} as a 17.4' large FR type II radio structure consisting of two extended lobes with a total flux density of only 87 mJy and a central compact core. The core is unresolved with the 5\\arcsec$\\times$5\\arcsec\\hspace{1mm} beam in the FIRST survey and has a flux density of 2.7 mJy. The structure is highly collinear and symmetric; the ratio of separation between the core and the brightest regions in the lobes is 1.08, and the misalignment angle is 1.3$\\degr$. The FIRST map reveals also compact hot spots in both lobes, which implies strong shocks and {\\sl in situ} reacceleration of relativistic particles, finally inflating the lobes. In order to check whether the observed lobes are or are not possibly connected by a bridge undetected in the NVSS because of the lack of short baselines, we made supplementary single-dish 21 cm observations using the Effelsberg 100 m radiotelescope, \\footnote{The Effelsberg 100m telescope is operated by the Max-Planck-Institut f\\\"{u}r Radioastronomie (Bonn, Germany)} as well as 50 cm observations with the Giant Metrewave Radio Telescope (GMRT).\\footnote{The GMRT is a national facility operated by the National Centre for Radio Astrophysics of the Tata Institute of Fundamental Research (Pune, India)} These observations, the final 1.4 GHz map of the investigated radio galaxy obtained from a combination of the NVSS and Effelsberg data, and the 619 MHz GMRT map are presented in \\S~2. The radio core position coincides perfectly with that of a galaxy with $R\\sim$19.65 mag and $B\\sim$21.82 mag at R.A.(J2000.0) = 14$^{\\rm h}$20$^{\\rm m}$23.80$^{\\rm s}$ and decl.(J2000.0) = $-$05\\degr45\\arcmin28.8\\arcsec, according to the magnitude calibration and astrometric position in the Digitized Sky Survey (DSS) data base. This DSS red magnitude suggests a rather distant galaxy whose photometric redshift estimate would be between 0.46 and 0.42, according to the Hubble diagrams for GRGs published by \\citet{schoen} and \\citet{lara}, respectively. However, the standard deviations in the apparent-magnitude distributions in those diagrams for a given redshift are about 1 mag. Therefore, supposing that the absolute magnitude $M_{\\rm R}$ of galaxies identified with very faint GRGs can be 1 $\\sigma$ lower than the mean value, we have estimated a value of $z\\approx 0.32$ for J1420$-$0545 (cf. \\citealt{mkj}). The independent photometry and optical spectroscopy of the host galaxy that confirm the above estimate are given in \\S~3, while some physical parameters and the dynamical age of the radio structure are estimated and discussed in \\S~4. ", "conclusions": "" }, "0808/0808.3744_arXiv.txt": { "abstract": "\\noindent{\\it The high-lights of ground-based very-high-energy (VHE, $E>100$~GeV) gamma-ray astronomy are reviewed. The summary covers both Galactic and extra-galactic sources. Implications for our understanding of the non-thermal Universe are discussed. Identified VHE sources include various types of supernova remnants (shell-type, mixed morphology, composite) including pulsar wind nebulae, and X-ray binary systems. A diverse population of VHE-emitting Galactic sources include regions of active star formation (young stellar associations), and massive molecular clouds. Different types of active galactic nuclei have been found to emit VHE gamma-rays: besides predominantly Blazar-type objects, a radio-galaxy and a flat-spectrum radio-quasar have been discovered. Finally, many (presumably Galactic) sources have no convincing counterpart and remain at this point unidentified. A total of at least 70 sources are currently known. The next generation of ground based gamma-ray instruments aims to cover the entire accessible energy range from as low as $\\approx 10$~GeV up to $10^5$ GeV and to improve the sensitivity by an order of magnitude in comparison with current instruments. } ", "introduction": "The highest energy photons known are produced in astrophysical processes involving even more energetic particles presumably accelerated through stochastic acceleration mechanisms as suggested initially by \\citet{1949PhRv...75.1169F}. Therefore, observations of VHE photons provide a direct view of the astrophysical accelerators of charged particles and allow to identify the individual sources of cosmic rays: VHE photons open our view to the ``accelerator sky''. The origin of the Galactic population of charged cosmic rays has remained since its discovery by V. Hess in 1912 until today a long-standing question of astro- and particle physics. A widely favored model on the origin of cosmic rays assumes that diffusive shock acceleration takes place in the expanding blast waves of supernova remnants converting $10-20~\\%$ of the kinetic energy into cosmic rays. Under this assumption, Galactic supernova remnants provide sufficient power ($\\approx 10^{41}$~ergs~s$^{-1}$) in order to balance the escape losses of Galactic cosmic rays as well as produce a power-law type distribution of particle energy that closely resembles the cosmic ray spectrum measured locally \\citep[see e.g.][]{1964ocr..book.....G,2006astro.ph..7109H}.\\\\ Observations of VHE-gamma-rays, mainly by imaging air Cherenkov telescopes in the last decade, have surpassed the anticipated detection of a few supernova remnants and have established a rich and diverse collection of VHE sources. Specifically, the currently active generation of imaging air Cherenkov telescopes (H.E.S.S.\\footnote{http://www.mpi-hd.mpg.de/hfm/HESS/HESS.html}, MAGIC\\footnote{http://wwwmagic.mppmu.mpg.de/}, and VERITAS\\footnote{http://veritas.sao.arizona.edu/}) have fulfilled and by far exceeded the expectations that were based upon the pioneering previous generation of experiments: the results obtained in the last years have shown that VHE emission is common to a variety of different source types - not only shell-type SNRs. \\\\ The current view of the source distribution in the VHE sky is shown in Figure~\\ref{fig:sky} where all known sources are displayed (status of September 2008). Note, the sensitivity achieved varies greatly across the sky. The best sensitivity is reached in the inner Galactic disk where a dedicated survey with the air Cherenkov telescopes of the H.E.S.S. experiment has been performed (see section~\\ref{subsect:imaging}).\\\\ The most remarkable feature of the gamma-ray sky is not evident in this picture: each source shown is an accelerator of particles up to and beyond TeV energies (``accelerator sky''). A similar conclusion can not be drawn when looking at source populations detected in other wavelength bands. This makes the VHE band a sensitive window to detect non-thermal particle accelerators. Future neutrino telescopes will almost certainly have the potential to detect some of the brighter VHE sources (and discover new sources that are optically thick for VHE gamma-rays). The detection of high energy neutrinos is experimentally a challenging task and requires tremendous efforts (see e.g. the Ice-CUBE neutrino telescope in the Antarctic ice). However, the observation of a neutrino source will ultimately demonstrate the presence of accelerated nuclei - largely model-independent. Different to the clear observational signature in the neutrino channel, VHE gamma-rays are sensitive to accelerated electrons (through inverse Compton scattering) as well as to accelerated nuclei (through neutral meson production and decay). The scope of this article is limited to observational highlights obtained with ground based instruments and does not aim at presenting a complete review of the field of VHE astrophysics. The paper is structured in the following way: In section~\\ref{section:obstech}, the experimental techniques of ground-based instruments (both imaging and non-imaging) are described before continuing with the census of today's VHE sky in section \\ref{section:census}. Section~\\ref{sec:secphy} provides a a short overview on VHE gamma-rays as probes of the interstellar medium including Lorenz-invariance violating effects related to structure of space-time at Planck-scales. Finally, in section~\\ref{section:future} this review is concluded with some comments on the future of the field. \\begin{figure} \\begin{center} \\includegraphics[angle=90,width=1.00\\linewidth]{vhesky.png} \\end{center} \\caption{\\label{fig:sky} All VHE sources as known in September 2008 are shown as circles in a Hammer-Aitoff projection of a Galactic coordinate system. The Galactic center is in the center of the picture. } \\end{figure} ", "conclusions": "The observational field of VHE gamma-ray astrophysics has been successfully driven in the last years by ground based imaging air Cherenkov telescopes. The ``high energy frontier'' of Astrophysics will be expanded by the results expected from the Fermi satellite. The all-sky sensitivity of the Fermi Large-Area-Telescope (LAT) will most likely lead to many interesting new discoveries and surprises. The inter-play of the Fermi-LAT detections from space and follow-up observations from ground will provide mutual benefits for the communities as well as certainly answer many questions as well as stimulate new ones. \\\\ While space-based observations of gamma-rays will be for a long time limited to the results obtained with Fermi (until maybe pair conversion telescopes will be deployed on the surface of the moon), ground based observatories will continue to be developed and constructed during the next decade(s). To my knowledge, besides Fermi, no further space-based gamma-ray telescope, operating in the GeV energy range, is planned. Consequently, ground based gamma-ray detection techniques will be the only available experimental approach to observe high energy gamma-rays after the termination of the Fermi mission.\\\\ The extensions of H.E.S.S. and MAGIC into phase II until 2009 will lower the energy threshold and improve existing sensitivity moderately. The next generation of ground based installations will become fully operational at the end of the next decade as envisaged in the AGIS\\footnote{Advanced gamma ray imaging system}\\citep{2007arXiv0709.0704K} and CTA\\footnote{Cherenkov telescope array} \\citep{2007AN....328..600H} projects. These installations aim at an improvement in sensitivity by a factor of 10 and a widened reach in energy. At that time, the new ground based instruments including larger ($>$km$^2$) installations like TenTen \\citep{2008NIMPA.588...48R} will explore new energy windows above $\\approx 10$~GeV as well as above 10~TeV. \\\\ However, there still remains a ``blind spot'': the transient sky above 10~GeV will neither be explored very well with Fermi (limited photon rate) nor with ground based instruments of the current generation (limited field of view and energy threshold). New installations like HAWC \\citep{2005AIPC..745..234S} will improve the situation in the energy range above a few 100~GeV. The lesson from the detection of fast transient events as seen from Cyg~X-1 and PKS~2155-304 however is that we are very likely missing the fast variability of XRBs and AGN. Proper coverage of these objects requires instruments with a large field of view and an energy threshold well below 100~GeV. \\label{section:future} {\\small" }, "0808/0808.3602_arXiv.txt": { "abstract": "The young system RX~J0529.3$+$1210 was initially identified as a single$-$lined spectroscopic binary. Using high$-$resolution infrared spectra, acquired with NIRSPEC on Keck II, we measured radial velocities for the secondary. The method of using the infrared regime to convert single$-$lined spectra into double$-$lined spectra, and derive the mass ratio for the binary system, has been successfully used for a number of young, low-mass binaries. For RX~J0529.3$+$1210, a long-period (462 days) and highly eccentric (0.88) binary system, we determine the mass ratio to be 0.78 $\\pm$ 0.05 using the infrared double-lined velocity data alone, and 0.73 $\\pm$ 0.23 combining visible light and infrared data in a full orbital solution. The large uncertainty in the latter is the result of the sparse sampling in the infrared and the high eccentricity: the stars do not have a large velocity separation during most of their $\\sim$1.3 year orbit. A mass ratio close to unity, consistent with the high end of the 1$\\sigma$ uncertainty for this mass ratio value, is inconsistent with the lack of a visible light detection of the secondary component. We outline several scenarios for a color difference in the two stars, such as one heavily spotted component, higher order multiplicity, or a unique evolutionary stage, favoring detection of only the primary star in visible light, even in a mass ratio $\\sim$1 system. However, the evidence points to a lower ratio. Although RX~J0529.3$+$1210 exhibits no excess at near-infrared wavelengths, a small 24~$\\mu$m excess is detected, consistent with circumbinary dust. The properties of this binary and its membership in $\\lambda$ Ori versus a new nearby stellar moving group at $\\sim$90~pc are discussed. We speculate on the origin of this unusual system and on the impact of such high eccentricity, the largest observed in a pre$-$main sequence double$-$lined system to date, on the potential for planet formation. ", "introduction": "Understanding the process of star formation requires reliable observations of fundamental stellar properties so that theoretical models can be tested. The copious population of young binary systems, however, complicates the problem. The binary fraction can be high \\citep{2000prpl.conf..703M}, and may depend on the density of the star forming region \\citep{2003ApJ...583..358B} and spectral type of the primary \\citep{2006ApJ...640L..63L}, underscoring the importance of their study. Fortuitously, binary stars serve two purposes. One, characterization of their frequency, separation distribution, and mass ratio distribution for a given star forming region (SFR) provides clues to the broad star forming properties (angular momentum, density, turbulence, etc.) of the parent molecular cloud. Two, the special class of very small separation spectroscopic binaries with periods sufficiently short to enable the measurement of the individual stellar velocities, and thus the system's mass ratio, are potential targets for the dynamical determination of individual component stellar masses (e.g., Steffen et al. 2001; Prato et al. 2002a; Boden et al. 2005; Stassun et al. 2007). Knowledge of absolute masses and observable properties, such as effective temperature and luminosity, plays a key role in the improvement of pre-main sequence (PMS) evolutionary models \\citep{2001ApJ...553..299P}. Once spectroscopic binaries have been identified, it is necessary to characterize their properties. It is often the case with low mass-ratio systems that they are identified as single$-$lined spectroscopic binaries when observed in visible light (Mazeh et al. 2002), in which case the large difference in flux between the primary and secondary at short wavelengths prevents detection of the secondary component and thus the measurement of the mass ratio for the system. In the Raleigh-Jeans regime, however, flux scales much less steeply as a function of mass. Thus, by observing single-lined spectroscopic binaries with infrared (IR) spectroscopy we are able to improve our chances of detecting the lower-mass secondary not seen in visible light. This technique was initially outlined in Prato et al. (2002b) and Mazeh et al. (2002, 2003). Our primary motivation for observing RX~J0529.3$+$1210 was to use this IR approach to determine the mass ratio of the system by converting it into a double-lined spectroscopic binary. Given that the secondary had not been detected in visible light, we anticipated a relatively low mass ratio for RX~J0529.3$+$1210. With the advent of the {\\it R\\\"ontgensatellit (ROSAT)} all-sky survey, a number of researchers began a search for X-ray sources with optical counterparts of brightness consistent with membership in nearby SFRs, motivated in part by the goal of detecting the ``post-T Tauri'' population postulated by Herbig (1978). In follow up observations of X-ray sources near Taurus, RX~J0529.3$+$1210 was identified variously as a PMS star (Neuhauser et al. 1997; Magazz\\`u et al. 1997) and a post-T Tauri star (Magazz\\`u et al. 1999); high-resolution visible light spectroscopy revealed its spectroscopic binary nature \\citep{neu97}. \\citet{tor02} undertook a seven year campaign to characterize the orbits of all spectroscopic binaries identified in the X-ray sample of \\citet{neu95} and \\citet{neu97}, including RX~J0529.3$+$1210. Table 1 summarizes the general properties of this system. The $\\sim$3 dozen high-resolution visible light spectra taken of this binary suggested the presence of a secondary star; however, a conclusive identification was not possible. A single-lined orbital solution was determined, although, probably owing to the extremely high eccentricity of the system the uncertainties are relatively large. This paper describes the results of using high resolution IR spectroscopy to observe RX~J0529.3$+$1210 and determine the component radial velocities. RX~J0529.3$+$1210 is the most eccentric pre$-$main sequence spectroscopic binary known to date. This provides a unique context in which to speculate on the formation of the system and on the impact of the stellar dynamics on potential planet formation. In \\S2 we briefly describe our observations and data reduction. Our analysis and results appear in \\S3. Section 4 provides a discussion, and \\S5 summarizes our findings. ", "conclusions": "\\subsection{Where and How Old is RX~J0529.3$+$1210?} The projected location of RX~J0529.3$+$1210 is associated with a high-density area in the CO maps of \\citet{dol01} for the $\\lambda$ Ori region. However, \\citet{dol01} find a mean radial velocity for the strong lithium sources identified in $\\lambda$ Ori of 24.5~km~s$^{-1}$, with a dispersion of only 2.3~km~s$^{-1}$. Our center-of-mass velocity for RX~J0529.3$+$1210 is 18.38~km~s$^{-1}\\pm0.30$, indistinguishable from that found by \\citet{tor02}. Thus, on the basis of radial velocities alone, it seems unlikely that the system is associated with $\\lambda$ Ori since RX~J0529.3$+$1210's center-of-mass velocity is inconsistent with that of $\\lambda$ Ori at the 3$\\sigma$ level. Using the 2MASS JHK magnitudes of RX~J0529.3$+$1210, an effective temperature (T$_{eff}$) for a K7/M0 of 3900~K \\citep{2003ApJ...593.1093L,bro06}, and the nominal distance to $\\lambda$ Ori (400~pc; Barrado y Navascu\\'es 2005) we have estimated the luminosity and placed the system on an H-R diagram. We find L $=$ 3.44 $\\pm$ 0.01 L$_{\\odot}$, yielding an age of $\\sim$0.1 Myr using the PMS tracks of \\citet{pal99}. Accounting for a companion star with luminosity equal to that of the primary results in an age of $\\sim$0.5~Myr. Using the tracks of \\citet{bar98}, or absolute K magnitude and the tracks of \\citet{2000A&A...358..593S}, we obtain similar age estimates. \\citet{bar05} estimates the age of $\\lambda$ Ori to be between 3 and 10 Myr, \\citet{bar07} find an age of 5~Myr, and \\citet{dol01} describe a spread in ages from 1 to 10 Myr. A $<<$1~Myr object is expected to be at least somewhat embedded in its natal cloud and associated with circumstellar material, yet near-IR colors from 2MASS magnitudes indicate zero extinction, no veiling is detected in the spectra, and the H$\\alpha$ emission line equivalent width (Table 1) is only 2\\AA. We consider two possible mechanisms for disk dissipation in this system, photoevaporation from nearby hot stars and the orbital dynamics of RX~J0529.3$+$1210 itself. The projected separation of the B8 star HD 36104 \\citep{dol01} from RX~J0529.3$+$1210 is $\\sim$1~pc if both are assumed to be at a distance of 400~pc. \\citet{dol01} discuss the surprising lack of evidence for accretion disks around the PMS stars in the central $\\lambda$ Ori cluster and suggest that photoevaporation and/or possibly a supernova event 1$-$2 Myr ago played a role in the dispersion of disks in the local young low-mass stellar population. RX~J0529.3$+$1210, located just north of the central cluster in the Barnard 30 dark cloud, could have experienced photoevaporation from the nearby B star at a young age, obliterating circumstellar material, although this is probably unlikely given the long survival time of proplyds in the Trapezium \\citep{1998AJ....115..263O}. Given the extremely high eccentricity of this binary, however, the action of the companion star may have precluded formation of, or dissipated any, circumstellar material. With an assumed primary star mass of 0.75 M$_{\\odot}$ and the mass function from Torres et al. (2002), we find a minimum secondary star mass of 0.40 M$_{\\odot}$, and thus a minimum total mass of 1.15 M$_{\\odot}$. In conjunction with the 461.89 day period, this yields a periastron separation of 0.15~AU and an apastron separation of 2.30~AU. It is unknown if orbital evolution may have occurred, or even whether it is possible that this unusual system formed by capture, but if RX~J0529.3$+$1210 is located in $\\lambda$ Ori with an age of $<$1~Myr, then it seems likely that the system formed in a similar configuration to its present one. Independent of the potentially harsh $\\lambda$ Ori environment and the orbital dynamics of RX~J0529.3$+$1210, the system manifests spectra of similar surface gravity to dwarf stars, as well as a relatively small lithium equivalent width, in comparison to classical T Tauri stars \\citep{1989AJ.....98.1444S,2005fost.book.....S}. The lithium and H$\\alpha$ equivalent widths of \\citet{dol01} for $\\lambda$ Ori average greater than twice what is found for RX~J0529.3$+$1210 and are consistent with the classical and weak-lined T-Tauri limits defined by \\citet{1998AJ....115..351M}. According to these limits, RX~J0529.3$+$1210 would be a post-T Tauri star and have one of the lowest lithium and H$\\alpha$ equivalent widths in the $\\lambda$ Ori region. These characteristics are inconsistent with an age of $<$1~Myr. Alternatively, RX~J0529.3$+$1210 could be an older, closer object. \\citet{mam07} describes a new candidate moving group, 32 Ori, consisting of a small cluster of X-ray bright, late type stars. The $\\sim$10 young stars identified in the group are located around 5$^h$ 20$^m$ to 5$^h$ 30$^m$ and $+6\\deg$, at the proposed distance of $\\sim$90~pc. RX~J0529.3$+$1210 is about 9~pc from the central clump of objects in 32 Ori and is one of the members used by Mamajek to define the common proper motion group (E. Mamajek 2008, private communication). Given the proper motion of RX~J0529.3$+$1210, pmRA $=$ 4.1 $\\pm$ 5.8 mas/yr and pmDec $=$ $-$30.7 $\\pm$ 5.8 mas/yr, compared to the proper motions of the stars $\\lambda$ Ori (pmRA $=$ 0.8 $\\pm$ 1.5, pmDec $=$ $-$2.3 $\\pm$ 1.5) and 32 Ori (pmRA $=$ 6.57 $\\pm$ 1.15, pmDec $=$ $-$32.45 $\\pm$ 0.48), evidence in support of membership in the 32~Ori group is compelling \\citep{1997A&A...323L..49P, 2004AJ....127.3043Z}. Again combining the 2MASS JHK magnitudes of RX~J0529.3$+$1210, an effective temperature of 3900~K, and a distance now of 90~pc, we find L $=$ 0.17 $\\pm$ 0.02 L$_{\\odot}$, giving an age for RX~J0529.3$+$1210 on the tracks of \\citet{pal99} of $\\sim$15$\\pm$5~Myr, consistent with the 25$\\pm$10~Myr age derived by \\citet{mam07} for the candidate group members. \\citet{mam07} lists a group radial velocity of 18~km~s$^{-1}$, in excellent agreement with our measured center-of-mass velocity (Table 2). Morales Calderon (2008, in prep) has observed the $\\lambda$ Ori region with the {\\it Spitzer} space telescope and finds a 3~$\\sigma$ detection of RX~J0529.3$+$1210 at 24~$\\mu$m with a 20~\\% excess above a 3900~K photosphere. For a star with L$=$0.17~L$_{\\odot}$, the equilibrium temperature of a black-body grain with peak emission at 24~$\\mu$m corresponds to a distance of 4.36~AU. A more realistic treatment of the dust grain distribution would necessarily take into account the additional flux from the secondary star, yielding a larger distance from the center-of-mass of the system to the putative dust. However, given the estimated apastron separation of 2.30~AU, a lower limit of 4.36~AU for the dust radius from the system center illustrates the plausibility of a circumbinary debris disk. Additional {\\it Spitzer} observations at shorter wavelengths have been taken and should reveal more information regarding the extent and location of the dust (Mamajek 2008, in preparation). The lack of J$-$H, H$-$K, and K$-$L excesses is hardly surprising. For black-body grains with a peak wavelength in the L band, the corresponding disk temperature occurs at a distance of $\\sim$0.1~AU, nearly coincident with the binary periastron. Stable dust in this system is most likely to be located in a circumbinary distribution. Furthermore, if the closer distance and therefore older age for this system implied by membership in 32 Ori are correct, then the presence of an evolved debris disk would not be unusual \\citep[e.g.,][]{tri08}. Finally, the fact that the RX~J0529.3$+$1210 secondary was possibly detected in our AO images strongly supports the $\\sim$90~pc distance. For the projected separation of 0.018\", determined using PSF fitting, and a distance of 90~pc, the corresponding distance in AU is 1.6$\\pm$0.54, not too different from our estimated apastron of 2.30~AU. For a distance of 400~pc the separation would be 7.2$\\pm$2.4 AU. Based on the currently available data, it therefore appears that RX~J0529.3$+$1210 is associated with the new 32 Ori group, not with $\\lambda$ Ori, and has an age of $\\sim$15~Myr. \\subsection{The Mass and Flux Ratios of RX~J0529.3$+$1210} \\citet{tor02} note that the appearance of the velocity correlation function for RX~J0529.3$+$1210 suggests the presence of at least one other star in the system. For a primary star mass of 0.75~M$_{\\odot}$, the mass function implies a minimum secondary star mass of 0.40~M$_{\\odot}$ and a minimum mass ratio of 0.53, consistent to within 1$\\sigma$ of the mass ratio found from the full orbital solution for the system, 0.73$\\pm$0.23. The approximate H-band flux ratio measured from the cross-correlation is 0.6$\\pm$0.1. The K-band flux ratio found in the best PSF fit to the January, 2008 AO images is 0.66$\\pm$0.18. For a M2.5 $+$ K7 pair, the spectral types which provided the best correlation (\\S 3), models of \\citet{pal99} and \\citet{bar98} imply an H-band flux ratio of 0.4$\\pm$0.1 based on components with a primary T$_{eff}=$3900~K and a secondary T$_{eff}=$3400~K \\citep{2003ApJ...593.1093L,joh66}. Taken together, these properties suggest a relatively red color for the secondary star, hindering detection in visible light. The IR primary and secondary velocities, presented in the Wilson plot (Figure 2), imply a range of mass ratios, from 0.73$-$0.83, well within the range we obtain by combining the IR and visible light data in a full orbital solution, $\\sim$0.50$-$0.93. Because the full solution includes phase information, it is generally more reliable, although for RX~J0529.3$+$1210 the sparse phase coverage during the epochs of largest velocity separation yields large uncertainties in the orbital solution. Clearly more data are merited. For the case of a mass ratio relatively close to unity in this system (i.e. 0.8$-$1.0), it is necessary to explain the lack of a visible light detection of the secondary star. In the following paragraphs we outline several scenarios that would account for this. A relatively red secondary and yet a mass ratio close to unity is possible if a third component is present, as first suggested by \\citet{tor02}. If the secondary star is actually a binary pair, with a very small separation and similar component masses, then the spectral types could be consistent with the best-fitting TODCOR templates. This inner binary pair would necessarily be in an orbit in a plane relatively perpendicular to our line of sight, such that the large velocity changes would not be obvious. This scenario would imply a near-unity mass ratio yet a smaller near-IR flux ratio, and a red enough secondary system to evade visible light detection. The drawback is that it requires a very specific geometry. Such a companion might also account for the 24~$\\mu$m excess. A heavily spotted secondary star could mimic a lower temperature source, resulting in a best fitting secondary M2.5 template, a mass ratio close to 1.0, and a flux ratio $<$1 in the near-IR. Again, the reddening effect of the cooler temperature in the spot covered areas would increase the difficulty of visible light detection. For two T$_{eff}=$3900~K stars, if one has 50\\% of its surface covered with a 2900~K spot \\citep{bou89}, the secondary/primary bolometric flux ratio would be $\\sim$0.7. For objects with ages between 10 and 30 Myr, some models \\citep{bar98,pal99} of PMS evolution indicate that targets in the mass range of the RX~J0529.3$+$1210 primary star, 0.6$-$0.8~M$_{\\odot}$, are located near the transition between the convective Hyashi track and the radiative Henyey track. As a result, models show a sharp turn in the mass tracks towards higher temperatures. A similar but slightly lower mass secondary star that has not yet turned this corner off of the Hyashi track, and indeed, objects with masses lower than $\\sim$0.5~M$_{\\odot}$ never will, may have a similar mass as the primary star but a much lower temperature. These models for the system behavior are highly speculative. The absence of an unambiguous secondary star detection in visible light suggests that the mass ratio is {\\it not} close to unity. A value in the 0.7 to 0.8 range is most likely based on data to date. \\subsection{Binary Formation Scenarios and the Potential for Planets} The formation of such a highly eccentric multiple is challenging to understand as a primordial event given current theories of star formation \\citep{2005fost.book.....S}. Figure 4 shows that pre$-$main sequence spectroscopic binaries are observed over a wide range of eccentricities. Although RX~J0529.3$+$1210 represents the maximum of that range, it does not stand out particularly from the eccentricity distribution. Thus, some form of dynamical evolution may have taken place in this system, but, if so, it is not distinguished by an outlying eccentricity. Possibilities for dynamical evolution range from an improbable capture event to disk excitation of stellar eccentricity \\citep{mat92}. If indeed this system is $>$10~Myr old, only fossil evidence, such as the detected 24~$\\mu$m excess, might remain of the primordial, presumably massive disk(s) responsible. In a variation of the interactions proposed by \\citet{rei00}, a three body dynamical encounter between an object from outside of the binary and a companion short period binary could have stimulated the eccentricity of the system and tightened the orbit of a close (secondary) pair (\\S 4.2). The eventual formation of a circumbinary (or circumtriple) debris disk could have been stimulated by dynamical disruption of the disks in the system in any one of these scenarios. The formation of planets in binary systems has been observationally supported in recent years and is of primary significance since a majority of stars have companions \\citep{egg07}. \\citet{cun07} have determined the strict criteria for classifying stable and unstable planetary orbits in binary systems, however, their analysis is restricted to circular orbits. Many of the extra-solar planets observed in binaries are in wide separation systems, where the mass ratio for the stellar components is near unity and the eccentricity of the system is low. \\citet{qui07} have modeled the formation of terrestrial planets around individual stars in binaries and find that, for periastron distances of $<$5~AU, such planet formation is restricted. Considering that the periastron distance in the RX~J0529.3$+$1210 system is only 0.15~AU, we conclude that the probability of even a low-mass circumstellar planet in RX~J0529.3$+$1210 is remote. The trend determined by \\citet{qui06} towards the formation of fewer circumbinary terrestrial planets in systems with apastron distances of $>$0.2~AU and non-zero eccentricities also bodes for poor circumbinary planetary stability around the stars in RX~J0529.3$+$1210. \\subsection{Improving our Understanding of the RX~J0529.3$+$1210 System: Future Work} Clearly it is imperative to improve the orbital solution for RX~J0529.3$+$1210, particularly the measurement of the mass ratio. Visible light and especially infrared data are critical to obtain, preferably at high precisions. RX~J0529.3$+$1210 passes through its next maximum velocity separation in late December, 2009. Densely sampled high signal-to-noise spectra taken during these epochs will yield a greatly improved precision for the orbital solution. If RX~J0529.3$+$1210 is located at a distance of only 90~pc, an apastron separation of 2.30~AU implies a maximum projected separation on the sky of $\\sim$0.03$''$. Much of the orbit of this system lies within reach of the diffraction limit of the 85~m baseline of the Keck Interferometer, 0.005$''$. Although the K$-$band magnitude of RX~J0529.3$+$1210, 9.2, is relatively faint for observations with this facility, planned improvements to the system may eventually enable resolved observations of this unusual binary, permitting the determination of individual component masses. The 90~pc distance may also be confirmed by measuring the parallax, which is slightly greater than 10 mas at this distance, and is at the achievable limit of current instrumentation." }, "0808/0808.2815_arXiv.txt": { "abstract": "We present the first detection of CS in the Antennae galaxies towards the NGC~4038 nucleus, as well as the first detections of two high-J (5-4 and 7-6) CS lines in the center of M~82. The CS(7-6) line in M~82 shows a profile that is surprisingly different to those of other low-J CS transitions we observed. This implies the presence of a separate, denser and warmer molecular gas component. The derived physical properties and the likely location of the CS(7-6) emission suggests an association with the supershell in the centre of M~82. ", "introduction": "\\label{sec:intro} The molecule CS is a good tracer of dense gas (n(H$_2$)$\\geq$ 10$^{5}$-10$^{7}$ cm$^{-3}$) in massive star-forming regions in our own Galaxy (\\citealt{Plum92, Plum97, Andr07}) and in nearby galaxies such as the Magellanic Clouds \\citep{Niko07}, M~51 and Maffei~2 \\citep{Pagl95}, NGC~253 \\citep{Maue89a, Maue89b, Mart05}, IC~342 and M~82 \\citep{Henk85, Maue89a, Maue89b}. Multi-line studies of CS are crucial for the determination of the average gas densities in galaxies since CS transitions have excitation thresholds ranging from 10$^{4}$-10$^{5}$ cm$^{-3}$ for the CS(2-1) line \\citep{Bron96} up to $\\sim 2\\times 10^{7}$ cm$^{-3}$ for the CS(7-6) transition \\citep{Plum92}. In this letter, we report in the center of M~82 the first detections of the CS(5-4) and CS(7-6) transitions and the detection of the CS(5-4) line towards the Antennae galaxies (NGC~4038). We have also re-observed lower-J CS lines in M~82. These two well-known sources were chosen because interferometric submillimeter/millimeter maps previously obtained have shown high concentrations of molecular gas. High resolution $^{12}$CO(1-0) maps indicate that the nearby (D = 13.8Mpc, see \\citealt{Savi04}) Antennae interacting galaxies are likely sites of rapid high mass star formation \\citep{Wils00}. M~82 represents an excellent example of a nearby (D=3.25 Mpc, see \\citealt{Dumk01}) starburst galaxy. Low-J molecular line studies (e.g. \\citealt{Fuen06, Seaq06, Mart06}) have determined its average gas physical parameters. However, the molecular emission from the M~82 nucleus appears to come from multiple gas components. In particular the detection of abundant CH$_{3}$OH \\citep{Mart06} and HCN \\citep{Brou93} indicate the presence of high density gas in the nucleus. Our detection of the CS(7-6) line not only clearly confirms the presence of very high density gas but its analysis (see Sect.~\\ref{sec:discu}) suggests that it is located in the expanding superbubble likely to be associated with supernoave remnants \\citep{Weis99, Will99, Yao06}. ", "conclusions": "\\label{sec:discu} The angular resolution of the CS(7-6) (beam size of 14\") does not allow us to resolve spatially the emitting region. However we can estimate its location from the radial velocity and the velocity width of this line by using the kinematic information provided by interferometric maps of other molecules. \\citet{Brou93} presented a 2\" resolution HCN(1-0) map of the south west part of M~82. From the channel map of the HCN(1-0) emission centered around 213.3 kms$^{-1}$($\\pm 40$kms$^{-1}$), it appears that the CS(7-6) emission may arise from the HCN(1-0) clump located at $\\alpha_{\\rm J2000}=09^h55^m52.2^s$ and $\\delta_{\\rm J2000}=69^\\circ40'46.1''$, within our beam. \\citet{Garc01, Garc02} presented high resolution interferometric HCO and H$^{13}$CO$^{+}$(1-0) maps of M~82. We re-analyzed the data and found that some of the clumps show narrow ($\\leq$ 60 kms$^{-1}$) emission, similar to that of our CS(7-6) line, confirming the likely location of the CS(7-6) emission at $\\approx$ (-5$''$, -2$''$) from the center. This is the location of the expanding molecular supershell \\citep{Weis99, Will99, Yao06}. The high density ($\\geq 10^{6}$cm$^{-3}$) and the high temperature ($\\approx$ 60-80K) may arise from the interaction between the expanding supershell and the ambient gas in M~82. \\begin{table*} \\caption{Observational parameters.}\\label{tab:obs} \\renewcommand{\\footnoterule}{} \\hspace*{-2cm} \\begin{tabular}{l c c c c c c c c c } \\hline Source & Line & $\\nu$ & Tsys & beam & $\\int$(T$_{mb}$ dv)& V$_{LSR}$ & $\\Delta$V$_{LSR}$ & T$_{peak}$ & rms\\\\ & & (GHz) & (K) & size (\") & (Kkms$^{-1}$) & (Kkms$^{-1}$) & (Kkms$^{-1}$) & (mK) & (mK)\\\\ \\hline M~82 & CS(2-1) & 97.980 & 281 & 25 & 13.3$\\pm$0.3 & 221.9$\\pm$2.7 & 225.7$\\pm$5.4 & 55.5 & 5.8\\\\ & CS(3-2)& 146.969 & 303 & 17 & 11.2$\\pm$0.3 & 219.7$\\pm$2.8 & 211.3$\\pm$6.4 & 50.1 & 4.6\\\\ & CS(4-3)& 195.954 & 1655 & 13 & 11.1$\\pm$1.3 & 161.0$\\pm$9.7 & 165.3$\\pm$23.2 & 63.4 & 25.8\\\\ & CS(5-4)& 244.936 & 437 & 20 & 4.2$\\pm$0.8 & 140.0$\\pm$38.0 & 211.2$\\pm$44.2 & 15.6 & 10.8\\\\ & CS(7-6)& 342.883 & 334 & 14 & 2.2$\\pm$0.2 & 213.8$\\pm$1.5 & 40.1$\\pm$2.8 & 52.5 & 14.1\\\\ \\hline NGC & CS(5-4)& 244.936 & 246 & 20 & 1.7$\\pm$0.1 & 1655.0$\\pm$4.4 & 98.7$\\pm$8.5 & 16.3 & 6.7\\\\ 4038&&&&&&&&&\\\\ \\hline \\end{tabular} \\end{table*} \\begin{figure} \\includegraphics[height=12cm]{fig1c.jpg} \\caption{Spectra of the CS(2-1), CS(3-2), CS(4-3), CS(5-4) and CS(7-6) lines from top to bottom, respectively, measured towards the center of M~82. The observed position corresponds to the center of M~82 ($\\alpha_{\\rm J2000}=09^h55^m51.9^s$ and $\\delta_{\\rm J2000}=+69^\\circ40'47''$). The CS(2-1) and the CS(4-3) spectra have been smoothed to a common velocity resolution of 6.1kms$^{-1}$ while the CS(3-2), CS(5-4) and CS(7-6) spectra have a velocity resolution of 8.2kms$^{-1}$, $\\approx$ 10kms$^{-1}$ and 6.8kms$^{-1}$, respectively. The velocity scale (x axis) is expressed in kms$^{-1}$ units while the temperature scale (y axis) is expressed in main beam temperature units (T$_{MB}$), converted from the antennae temperature via the main beam efficiencies listed in Sect. ~\\ref{sec:obs}. The solid black lines superimposed onto each spectrum represents the Gaussian fit while the vertical black line marks the systemic M~82 V$_{LSR}$. }\\label{fig:1} \\end{figure} \\begin{figure} \\includegraphics[height=4cm]{fig2.jpg} \\caption{Spectrum of the CS(5-4) line towards NGC~4038. The observed position corresponds to the NGC~4038 nucleus ($\\alpha_{\\rm J2000}=12^h01^m52.8^s$ and $\\delta_{\\rm J2000}=-18^\\circ52'05''$). The spectrum has been smoothed to a velocity resolution of 2.3 kms$^{-1}$. The solid black line superimposed on the spectrum represents the Gaussian fit while the vertical black line marks the systemic NGC~4038 V$_{LSR}$.}\\label{fig:2} \\end{figure} \\begin{figure} \\includegraphics[height=4cm]{fig3.jpg} \\caption{Rotation diagram derived from the CS lines towards M~82. The black lines represent the linear regression fits for two gas components while the dashed line corresponds to one single gas component. The black solid squares show the data with error bars. These error bars usually represent the main errors in rotational diagrams which actually correspond to those of the integrated intensities (order of 10-20\\%). This population diagram has been corrected for beam dilution effects assuming a source size of 10$''$ for the center of M~82. This value is estimated from the interferometic data presented in \\citet{Brou93} and \\citet{Garc01}.}\\label{fig:3} \\end{figure} \\vspace*{1cm} In summary, we have presented the first detection of a high density tracer (CS) in the Antennae galaxies (NGC~4038 nucleus) as well as the first detections of two high-J transitions of CS in the center of M~82. We find that multiple molecular gas components in M~82 are necessary to explain the observed line intensities. In particular, while low-J CS lines seem to arise from relatively low density gas ($\\sim 10^{5}$cm$^{-3}$), the molecular gas traced by the CS(7-6) line must be dense ($\\sim 10^{7}$cm$^{-3}$) and warm ($\\sim$ 70K) and appears to be associated with the expanding supershell in M~82. The high density and temperature may be due to the interaction between the expanding supershell and the ambient gas. Similar multi-line studies for the Antennae galaxies (NGC~4038/39) are necessary in order to determine the origin of its high density molecular gas component traced by the CS(5-4) line. In both sources, the CS column densities are consistent with model predictions of gas undergoing high-mass star formation \\citep{Baye08}. Similar extended multi-transition multi-molecule studies performed on ultra-luminous infra-red galaxies by \\citet{Grev06b} and \\citet{Baan08} need to be carried out on nearby sources. In fact, \\citet{Mueh07} presented a detailed study of the para-H$_{2}$CO in M~82; however this molecule does not trace the dense gas component. Thus, in order to compare the physical and chemical properties of the very dense (extragalactic) gas between ULIRGs and nearby sources, CS observations (from 2-1 to 7-6) of ULIRGs should be performed." }, "0808/0808.0398_arXiv.txt": { "abstract": "{} {The main purpose of this work is to improve the existing knowledge about the most powerful engines in the Universe -- quasars. Although a lot is already known, we still have only a vague idea how these engines work exactly, why they behave as they do, and what the relation is between their evolution and the evolution of their harboring galaxy.} {Methods we used are based on optical spectroscopy of visually bright quasars, many of which have recently been discovered as X-ray sources, but eventually missed in color-selected surveys. The spectra typically cover the 4200--7000 \\AA\\AA\\ region, allowing measurements of the characteristics of the hydrogen lines, the FeII contribution, and other lines of interest.} {We present accurate redshift estimates and Seyfert type classification of the objects. We also show that the contribution of the host galaxy to the optical continuum is non-negligible in many cases, as is the intrinsic AGN absorption. Consequences of not correcting for those factors when estimating different quasar parameters are discussed. We also find some evidence of a non-unity slope in the relation between the internal extinction based on the Balmer decrement and the one on the optical continuum slope, implying, if further confirmed, the intriguing possibility that some absorbing material might actually be located $between$ the continuum source and the broad-line region.} {} ", "introduction": "The generally accepted paradigm about the nature of quasars (or active galactic nuclei; AGN) invokes a supermassive black hole, surrounded by an accretion disk, where most of the emitted energy is generated. In the vicinities of this disk, there must be a region, producing the broad emission lines (BLR), traditionally presented as a system of clouds, moving in the gravitational field of the black hole. Farther out from the center should be a region producing narrow lines (NLR), which may be the only visible optical lines in the spectra, if the central region is hidden from the observer behind a thick torus, as the unification models explain the AGN type I/II differences (Antonucci, 1993, for a review). Studying the narrow lines is important, since they are fairly sensitive to the shape of the ionizing continuum, and especially to the unobserved EUV range, which, if restored properly, may help to test the general paradigm or to model the type of accretion disk (e.g. standard thin disk, ADAF, etc.). One of the least clear points in this picture so far appears to be the exact geometry (e.g. flat, spherical, conical), structure (system of clouds, irradiated accretion disk, etc.), and kinematics (inflow, outflow, Keplerian rotation) of the BLR matter. The BLR should play an important role in feeding the quasar; therefore, knowledge of its nature may reveal why some galactic centers are active and others are not, in case a supermassive black hole is likely to be present in both types (Magorrian et al. 1998). The best way to probe the broad-line region might be through its optical emission-line spectrum. Although a single spectrum is certainly not sufficient to answer all the questions, studying the line profiles, ratios, and other properties in many objects seems to be one way of sheding some light on the problem. A promising new approach to systematizing quasar properties and differentiating between quasars is based on principle component analysis (PCA), where the quasar diversity can be expressed as a linear combination of a few eigenvectors of the correlation matrix, constructed from different measured quantities of a sample of quasars. Many of these measures are based on the optical spectra (Boroson \\& Green 1992; Sulentic et al. 2000), such as the width of the broad lines, equivalent width ratios. It is proposed that the main eigenvector (EV1), i.e. the main driver of the quasar diversity, is the accretion rate ratio (Marziani et al. 2001), one of the defining characteristics of the AGN. On the other hand, another important characteristic of the central engine, the black hole mass, is often calibrated through empirical relations, like the ``width of the broad lines, FWHM -- the monochromatic continuum luminosity, $\\lambda L_{\\lambda5100}\\AA$'', (Kaspi et al. 2005), ``the black hole mass -- bulge luminosity or bulge velocity dispersion'' (Magorrian et al. 1998; Ferrarese \\& Merritt, 2000), etc., since the only direct method known, the reverberation mapping (see Peterson \\& Horne, 2004, for a review), is observationally very expensive. It is clear that a large sample of quasars with relatively good spectra is needed for better understanding the importance of EV1 correlations for the quasar physics and to better calibrate different quasar empirical relations. Nowadays we see that many bright broad-line quasars have been missed in color-selected samples (e.g. Palomar--Green, Schmidt \\& Green, 1983), mainly because of their reddened colors, due to significant host galaxy contribution and/or intrinsic absorption. Being mostly radio-quiet, many of those objects have only recently been discovered and identified as quasars through X-ray surveys. Studying such objects in detail is especially important for preventing introduction of different biases into our knowledge about the overall AGN population, its properties, and evolution. Furthermore, these partially obscured objects (reddened S1, S1.8, S1.9 types, etc.) may provide an important link between the ``pure'' type I AGN and the obscured population (type II / Seyfert 2), especially taking into account the building evidence that the latter may comprise a significant part of the entire AGN population (Zakamska et al. 2003). This paper should be considered as a part of a larger continuing effort to obtain spectra of relatively good quality for most of the nearby, visually bright quasars, including many discovered in recent years. In a subsequent paper, we will concentrate on the connection between the quasar host galaxies and different AGN characteristics. The paper is organized as follows. In the next two chapters we describe the observations and present the main results, including individual notes on the objects. Discussion is presented in Sect. 4 and the summary, in Sect. 5. \\begin{table*} \\caption{Log of observations.} \\label{table:1} \\centering \\begin{tabular}{lcccccccccc} \\noalign{\\smallskip} \\hline\\hline Object & RA J2000 & Dec J2000 & z & $\\sigma$(z) & A$_{\\rm V}$ & Date & Instr. & Exp. & S/N\\\\ \\hline 2MASXJ005050+3536& 00 50 51 & +35 36 43 & 0.0585 & 0.0004 & 0.14 & 08.09.2004 & Rz 2.0 & 2700 & 36\\\\ MCG+08.17.060 & 09 13 46 & +47 42 00 & 0.0524 & 0.0005 & 0.05 & 19.03.2004 & Rz 2.0 & 2x2400 & 10\\\\ RXS J11401+4115 & 11 40 03 & +41 15 05 & 0.0717 & 0.0002 & 0.06 & 22.06.2005 & Sk 1.3 & 2x1200 & 12\\\\ PG 1211+143 & 12 14 18 & +14 03 13 & 0.0815 & 0.0003 & 0.11 & 23.06.2005 & Sk 1.3 & 2x1200 & 27\\\\ RXS J12308+0115 & 12 30 50 & +01 15 21 & 0.1183 & 0.0005 & 0.06 & 24.06.2005 & Sk 1.3 & 1200 & 25\\\\ RXS J16312+0955 & 16 31 16 & +09 55 58 & 0.0917 & 0.0005 & 0.18 & 17.08.2004 & Sk 1.3 & 2x2700 & 35\\\\ RXS J17233+3630 & 17 23 25 & +36 30 25 & 0.0400 & 0.0003 & 0.16 & 29.06.2005 & Rz 2.0 & 1800 & 25\\\\ IRAS 18423+2201 & 18 44 30 & +22 04 28 & 0.0464 & 0.0004 & 0.69 & 11.08.2005 & Sk 1.3 & 2x2400 & 15\\\\ NPM1G+27.0587 & 18 53 04 & +27 50 28 & 0.0620 & 0.0004 & 0.49 & 10.08.2005 & Sk 1.3 & 2x1800 & 31\\\\ & & & & & & 04.05.2005 & Rz 2.0 & 1800 & 29\\\\ RXS J20440+2833 & 20 44 04 & +28 33 09 & 0.0492 & 0.0005 & 1.03 & 28.08.2005 & Sk 1.3 & 2400 & 38\\\\ & & & & & & 30.06.2005 & Rz 2.0 & 1800 & 33\\\\ NPM1G$-$05.0589 & 21 03 38 &$-$04 55 40& 0.0575 & 0.0002 & 0.27 & 18.08.2004 & Sk 1.3 & 2x2700 & 31\\\\ RXS J21240$-$0021& 21 24 02 &$-$00 21 58& 0.0617 & 0.0003 & 0.15 & 17.08.2004 & Sk 1.3 & 2x2700 & 36\\\\ NPM1G+24.0470 & 21 39 41 & +24 24 18 & 0.0390 & 0.0005 & 0.23 & 10.08.2005 & Sk 1.3 & 2400 & 26\\\\ RXS J21592+0952 & 21 59 12 & +09 52 42 & 0.1003 & 0.0002 & 0.21 & 19.08.2004 & Sk 1.3 & 2x2700 & 39\\\\ RXS J22027$-$1304& 22 02 45 &$-$13 04 53& 0.0391 & 0.0003 & 0.14 & 28.08.2005 & Sk 1.3 & 2x2400 & 22\\\\ RXS J22160+1107 & 22 16 04 & +11 07 26 & 0.0514 & 0.0005 & 0.21 & 10.08.2005 & Sk 1.3 & 2400 & 16\\\\ RXS J22287+3335 & 22 28 46 & +33 35 08 & 0.0906 & 0.0004 & 0.29 & 11.08.2005 & Sk 1.3 & 2x2400 & 35\\\\ NPM1G$-$04.0637 & 22 53 11 &$-$04 08 49& 0.0257 & 0.0002 & 0.14 & 11.08.2005 & Sk 1.3 & 2x2400 & 24\\\\ \\hline\\hline \\end{tabular} \\end{table*} ", "conclusions": "\\subsection{Line profiles} While fitting the line profiles, we did not follow a multi-Gaussian/Lorentzian approach, because two or three, say Gaussians, fitting the broad profile can hardly have any physical meaning (in the sense that they can hardly represent different physical regions). Furthermore, each complex profile can be fitted to the desired level of accuracy with a large enough number of functions; however, the fit is not necessarily meaningful, not even unique in terms of degeneracy. We have to mention, though, that multi-Gaussian fits are often used by other authors, mainly for the sake of a better description of the profile (Warner et al. 2003; Popovic et al. 2004; see also Bachev et al. 2004, and the discussions there). The impression that the narrower profiles are better described by a Lorentzian, not Gaussian, (Sulentic et al. 2000; V\\'{e}ron-Cetty et al. 2001), however, may have no physical meaning, since it is not clear to what extent FWHM measures physically meaningful velocity-related quantity in a complex profile. A more adequate measure would be the second moment of the profile (the velocity dispersion), which is infinity for a Lorentzian. Even if this were not the case (say for a modified Lorentzian), a Lorentzian-type profile typically has higher dispersion than a Gaussian of the same FWHM. Therefore, it would be biased to conclude that from two profiles of the same velocity dispersion, the narrower one (in terms of FWHM!) resembles a Lorentzian more than a Gaussian. \\subsection{The obscured AGN population} Finding more and more rather bright (i.e. 14--15 magnitude) AGN shows clearly that color-based quasar surveys, like PG for instance, are not complete. This is obviously so, because the quasars are very different, not only in terms of emission-line, radio, X-ray properties, but also in terms of the optical continuum slopes, which ultimately define the colors. The intrinsic quasar continuum obviously may not only depend on the exact accretion solution type and accretion parameters, such as black hole mass, accretion rate, accretion disk inner and outer radii, etc. The continuum shape and level can be severely altered by the host galaxy contribution and the internal absorption, and we found evidence of both in the sample we studied. The reddened continuum slopes can eventually explain why such bright object are not discovered earlier as AGN. On the other hand, a correct assessment of the intrinsic $\\lambda L_{5100}$ is tightly linked to the estimates of the bolometric luminosity and the central mass through empirical relations (the often used $L_{\\rm bol}\\simeq 9 \\lambda L_{5100}$, see also Sect. 3.2). It also affects other characteristics, like optical-to-X-ray slopes, radio-loudness (e.g. Ho, 2002), line equivalent widths, etc. Therefore, one needs to know the intrinsic quasar continuum for accurately computing the accretion parameters and exploring correlations like EV1. For instance, large quasar samples with well-studied spectral properties (e.g. Boroson \\& Green, 1992; Marziani et al. 2003) rarely make an attempt to assess the real continuum level through correcting for the starlight and the intrinsic absorption, presuming that their contribution is small. It may indeed be so for some bright quasars, but as we see here (see also Dietrich et al, 2005), this is not always the case. Furthermore, dangerous biases may be introduced -- the host galaxy will contribute to the continuum mostly for a lower-luminosity AGN, and the internal absorption may be connected with the orientation if associated with a thick torus, and the broad-line widths if the torus is coplanar to a flattened BLR. It should be easy to correct for the host galaxy in a spectrum of relatively good quality. Unfortunately, the host galaxy contribution depends largely on the slit size, seeing, etc. and therefore may differ significantly from observation to observation and from instrument to instrument. This also can have consequences for the reverberation-mapping studies, in the sense that a variable starlight contribution can at least add some extra noise in the cross-correlation function between the continuum and the lines. \\subsection{Where is dust located?} In this paper we did not make any attempt to correct the continuum for the intrinsic absorption; nevertheless, the latter is rather apparent for objects like NPM1G-04.063, where quasar continuum around H$\\beta$ appears to be totally absorbed. One may make an attempt to take this absorption into account either by measuring the continuum slope or by comparing the Balmer decrement with the theoretical expectations (i.e. case ``B'' recombination). Both approaches may lead to incorrect results, however. The slope can, of course, vary for a number of reasons and the dust reddening is only one of them. On the other hand, the Balmer decrement can also be misleading since we do not know exactly where the absorbing material is. If it is uniformly spread outside the NLR, then such a correction can in principle be performed, and the absorption will presumably equally affect the NLR, the BLR, and the central continuum source. Different Balmer decrements for the broad and the narrow lines, as seen in some of the objects in our sample, suggest this may not always be the case. If the absorbing material is at least outside the BLR, then the broad line decrement could be used to assess the unabsorbed continuum level and shape. Unfortunately, we cannot rule out the presence of absorption inside the BLR inner radius (or mixed with the BLR matter), in which case the continuum will be affected, but indications for this might not be seen from the emission line ratios. An important clue to solving the problem of the dust location can be provided by the relation between the optical continuum slope and the Balmer decrement. We found a correlation between the slope and the decrement absorption indices (Fig. 5). One is to note, however, that the wide range of both -- Balmer decrements and continuum slopes observed in our rather small sample is not typical. For instance, Dong et al. (2008) report a much narrower range of Balmer decrements for a large sample of SDSS quasars -- typically between $-0.5$ and 1, and slopes -- between 1.2 and 2.2 (both expressed in terms of $A_{\\rm V}$). Their sample is, however, blue-selected and is obviously biased against the obscured AGN population. This may explain why Dong et al. (2008) find no correlation between the decrements and the slopes in their sample. If further confirmed, the relation $A_{\\rm V}^{\\rm cont}\\simeq 2 A_{\\rm V}^{\\rm BLR}$ (Sect. 3.4) suggests that the reddening on average affects the center more than the BLR. Even adopting a different value for the intrinsic continuum slope (like the often assumed $a_{\\rm \\nu}^{\\rm intr}\\simeq 1$) only offsets the relation between the Balmer decrement and the continuum slope, but does not change the slope of the relation. Adopting a different value for the intrinsic Balmer decrement (e.g. 2.8 instead of 3.1) has a similar effect. One way to explain this result is to assume that the dust is associated with the BLR matter and the reddening affects the continuum coming from inside more than the BLR emission itself. Another explanation would be that the dust is much farther outside the BLR (probably associated with a torus), and we see the active nucleus just above the torus edge. As a result, the central continuum can appear heavily absorbed, but the BLR light is not that affected, taking their different spatial dimensions into account. If the latter explanation turns out to be the correct one, studying the line profiles of such partially obscured objects can be important for probing the exact BLR geometry and kinematics, since knowing the torus opening angle, the inclination can be estimated for these objects." }, "0808/0808.1161_arXiv.txt": { "abstract": "In order to quantitatively test the ability of averaged inhomogeneous cosmologies to correctly describe observations of the large scale properties of the Universe, we introduce a smoothed template metric corresponding to a constant spatial curvature model at any time, but with an evolving curvature parameter. This metric is used to compute quantities along an approximate effective lightcone of the averaged model of the Universe. As opposed to the standard Friedmann model, we parameterize this template metric by exact scaling properties of an averaged inhomogeneous cosmology, and we also motivate this form of the metric by results on a geometrical smoothing of inhomogeneous cosmological hypersurfaces. The purpose of the paper is not to demonstrate that the backreaction effect is actually responsible for the Dark Energy phenomenon by explicitly calculating the effect from a local model of the geometry and the distribution of matter, but rather to propose a way to deal with observations in the backreaction context, and to understand what kind of generic properties have to hold in order for a backreaction model to explain the observed features of the Universe on large scales. We test our hypothesis for the template metric against supernova data and the position of the CMB peaks, and infer the goodness--of--fit and parameter uncertainties. We find that averaged inhomogeneous models can reproduce the observations without requiring an additional Dark Energy component (though a volume acceleration is still needed), and that current data do not disfavour our main assumption on the effective lightcone structure. We also show that the experimental uncertainties on the angular diameter distance and the Hubble parameter from Baryon Acoustic Oscillations measurements -- forseen in future surveys like the proposed EUCLID satellite project -- are sufficiently small to distinguish between a FLRW template geometry and the template geometry with consistently evolving curvature. \\\\ \\\\ {\\bf Keywords:} dark energy theory; CMBR experiments; supernova type Ia ", "introduction": "\\label{sec:intro} On the very large scales the Universe appears to be close to a homogeneous and isotropic state. This is usually modeled by a locally isotropic and hence homogeneous solution of Einstein's equations, namely the standard Friedmann--Lema\\^\\i tre--Robertson--Walker (FLRW) metric. The observations of tiny temperature fluctuations of the Cosmic Microwave Background (CMB) radiation suggest that in the Early Universe deviations from this `background cosmology' were very small, thus motivating the use of linear perturbation theory about the FLRW solution. The experimental confirmation of the predicted baryon acoustic oscillations in the CMB power spectrum \\cite{debernardis,spergeletal}, as well as the distribution of galaxies and clusters on the large scales \\cite{lahav} provide a certain body of evidence in favor of the `concordance model' which results from this perturbative approach (see, however, \\cite{marraPN}, \\cite{huntsarkar}). This standard scenario relies on the assumption that the FLRW cosmology correctly describes the `background cosmology', i.e. the averaged inhomogeneous Universe, at all times. Even though this hypothesis may be valid in the Early Universe, does it continue to hold even when the Universe becomes more and more structured at late times? Answering such a question has become even more important in the light of the still unexplained Dark Energy phenomenon in the context of the FLRW paradigm. The luminosity distance measurements to type Ia supernova (SN-Ia) standard candles, when analyzed within the framework of the FLRW Universe, provide strong evidence for a missing component characterized by a negative pressure which, by inducing an accelerated phase of expansion, would be responsible for the observed dimming of far distant SN Ia (\\cite{Perlmutter,Riess}). The simplest scenario to account for these observations is a positive cosmological constant in Einstein's equations. This is often assumed to describe the energy contribution of quantum vacuum fluctuations. Nevertheless, because of the huge discrepancy between the particle physics expected value and the observed one, several alternative scenarios have been investigated. For instance phenomenological models such as a late time slow rolling scalar field (see reviews \\cite{copeland,pilar}) or the Chaplygin gas \\cite{chaplygin} have been proposed to describe this Dark Energy component. Alternatively, there have been several proposals to account for these effects through modification of the laws of gravitation (e.g. braneworlds \\cite{maartens:brane}, scalar--tensor gravity \\cite{esposito}, higher--order gravitational theories \\cite{riccilagrangians,capo}, AWE \\cite{AWE1,AWE2}). Recently, a third alternative has been considered \\cite{rasanen:darkenergy,kolbetal} that aims at explaining Dark Energy as an effect caused by inhomogeneities. However, most of the approaches which include the effect of inhomogeneities still rely on the postulate that the FLRW solution reliably describes the effective (average) evolution of an inhomogeneous cosmology. For instance this is the case of inhomogeneous universe models in which distances are computed using perturbation theory about a FLRW background \\cite{ruth:luminosity,vanderveldetal}. Other efforts abandon the FLRW model and instead restrict inhomogeneities by strong symmetries, employing exact solutions to Einstein's field equations like the Lema\\^\\i tre--Tolman--Bondi (LTB) metric \\cite{celerier1, rasanen:LTB, nambu, LTBgron, LTBluminosity5, LTBluminosity1, LTBluminosity2, LTBluminosity4, alnes2, alnes3, singh1,Bolej,sussman2}. In this work we shall test a different approach and exploit the key--insight that the (large--scale) kinematics of a homogeneous--isotropic {\\em state} does not necessarily follow the kinematics of a homogeneous--isotropic {\\em solution}, especially at late epochs characterized by the presence of large matter inhomogeneities. Indeed, the analysis of backreaction effects due to inhomogeneities suggests that there is a wider class of (large--scale) homogeneous--(almost--)isotropic cosmological models, while smaller scales feature strong inhomogeneities and anisotropies that both are known to exist. In such a case it is natural to ask whether the emergence of Dark Energy can result from the breakdown of the underlying assumption associated with the FLRW cosmology. If no assumption on the nature of the inhomogeneities is made, i.e. if we do not restrict them to be small deviations from a FLRW background or obeying strong symmetry restrictions, we can still look at effective (average) properties of Einstein's equations. In the simplest case such a programme can be realized by foliating spacetime into flow--orthogonal hypersurfaces, restricting the matter model to `dust', and spatially averaging the scalar parts of Einstein's equations with respect to a collection of free--falling observers (generalized fundamental observers). Whereas such a formalism and the dynamical equations that govern the behavior of the averaged inhomogeneous universe model are well--established \\cite{buchert:review}, the explicit geometry, which lies at the basis of how we measure distances, is left unspecified. Here we suggest, as a next step, to complement the general kinematical properties of an averaged universe model with an explicit form of a {\\em template metric} that retains the main properties of the standard model of cosmology, such as its isotropy and homogeneity on large scales, but allows for structuring on small scales. The shift in emphasis is from postulating a {\\em strong cosmological principle} that assumes local isotropy about every point and hence homogeneity on all scales, to a {\\em weak cosmological principle} that only assumes (quasi--) isotropy and homogeneity on the largest observable scales. In this context, we retain by assumption the usual description of the Universe at early times, up to decoupling. However, at late times we will have to modify this description. To summarize let us list our main goals and assumptions. The purpose of this paper is to study the influence of a geometrical effect induced by the coupling between backreaction and averaged spatial curvature, on top of the well--studied kinematical effect of backreaction on the evolution of the effective volume scale factor \\cite{buchert:review}. To realize this study, we propose an ansatz for the effective metric of the large-scale homogeneous model that is motivated by previous results on the smoothing of Riemannian metrics by the Ricci flow. This is used to define an effective background on which the photons propagate; such a background can be considered as a first refinement of the usual FLRW background geometry. In order to get an insight into the effects associated with our prescription, we consider a specific example of backreaction, namely a power law of the effective scale factor. Here we want to stress the fact that our aim is not to show that backreaction effects can be fully responsible for the Dark Energy phenomenon. Instead, we address the converse problem: what is necessary and what kind of generic effects are expected for a backreaction model to be consistent with the cosmological observations? And more specifically, we are interested in understanding what specificities can allow to distinguish between FLRW and averaged models. The central concept underlying our investigation is that, although the 3--Ricci curvature distribution of an inhomogeneous cosmological slice can be smoothed {\\em at any time} into a constant curvature, the dynamical evolution of the averaged curvature can differ from the evolution of a constant--curvature (homogeneous) model. This deviation has recently been quantified in the framework of perturbation theory \\cite{lischwarz,lischwarz1,lischwarz2} (see also \\cite{Ian} for an estimation in the conformal Newtonian gauge), and since we have arguments why a non--perturbative treatment is necessary for the effects of interest, we shall consider the dominant perturbative mode within a general class of scaling solutions to a backreaction--driven cosmology. Of course, the scaling solutions cannot be expected to fully represent the realistic backreaction effect throughout the whole history of the Universe, but it is considered here for reasons of clarity and simplicity to illustrate the kind of effects expected from the non-trivial geometry, in analogy to studies using parameterizations of the equation of state for Dark Energy. The paper is organized as follows. We introduce the backreaction context, the key--equations and free parameters of an averaged cosmological model in Section~2. In Section~3 we develop the ansatz for the metric, which is inspired by the study of the Ricci flow deformation of three--dimensional Riemannian initial data sets, and we use this effective metric to compute quantities along an approximate past lightcone, in particular the luminosity distances of cosmological objects. This template metric can be considered as an improvement over the FLRW ansatz that considers that photons follow the null geodesics of a locally homogeneous and isotropic metric. In Section~4 we discuss the constraints on the model parameters as inferred from SN Ia data and the multipoles of the CMB peaks and dips. In particular we calculate the cosmic distances in the template metric and determine the `best-fit' models with the help of exact scaling solutions to the backreaction problem, which include the leading perturbative mode. We conclude this section with a discussion on possible tests of our main assumptions. Finally, we summarize the results of the paper and present an outlook in Section~5. ", "conclusions": "In this paper we have addressed the problem of comparing averaged inhomogeneous cosmologies with observational data by proposing to fit the observations with the help of an improved template metric, whose form is compatible with the kinematical integral properties of a general averaged model. This template metric has been motivated by the fact that the averaged curvature of space--like hypersurfaces of the space--time foliation is not expected to be constant in time. Indeed, the cosmological principle only requires the {\\it spacelike} quantities to be averaged, but does not impose anything on the evolution of these quantities. In other words, the FLRW universe models, which obey a strict cosmological principle, are a very particular subclass of models respecting a weaker cosmological principle presented in this paper. We consider the modified template metric as a first approximation tool for interpreting observations in a Universe that appears homogeneous {\\it on large scales}, but in which the backreaction effect cannot be neglected. That means that the proper effective lightcone along which the cosmological observations are made cannot be simply approximated by a FLRW lightcone. It is indeed important to notice that this template metric has only been introduced to compute quantities on the lightcone.\\\\ Thanks to this prescription for the lightcone, we then deduced constraints on the particular class of scaling backreaction, using the supernov{\\ae} luminosity/redshift distribution, and the positions of the peaks in the CMB spectrum. We found that the non--trivial geometry of the lightcone induces a slight change in the constraints, with respect to the same models in a FLRW geometry, leading to models compatible with the data for higher values of $\\Omega_{m}^{\\now\\CD}$, that is to say with a smaller amount of backreaction. This is particularly true for the leading perturbative mode ($n=-1$). One should note that the model presented in this paper still needs an acceleration of the effective volume scale factor to reproduce the data. Recent results based on the the Lema\\^\\i tre--Tolman--Bondi (LTB) solution (see \\cite{Bolej} and references therein) show that, on the contrary, one can fit the data without an acceleration of the volume scale factor. The two results do not necessarily disagree. Indeed, in the LTB model, one fits the data with an inhomogeneous metric having a functional degree of freedom ($t_{b}(r)$ in \\cite{Bolej}), and then, one averages the best--fit model to find that there is no acceleration of the effective volume scale factor. In this paper, we have first introduced an effective homogeneous model that we fitted to the data, and instead of having a functional dependence in the effective metric, we have made the assumption that backreaction features a scaling behavior with the scale factor. The comparison between the two approaches could be done once more realistic inhomogeneous metrics are at hand, and a more realistic behavior of the backreaction effect could be implemented. Volume acceleration might be a consequence of over--evolving the backreaction at early times by a strict scaling ansatz. One important feature of our results is that the cosmic history and the distances are strongly affected by the introduction of a non--FLRW geometry for the past lightcone, even if the constraints are only slightly different in the $(\\Omega_{m}^{\\CD},n)$ plane. Finally, we discussed two ways of testing the assumption (\\ref{eq:defkappa}) that links the geometry of the lightcone with the kinematical properties of the model. On the one hand, we found that the supernov{\\ae} data are fully compatible with the assumption that the averaged curvature felt by photons along the past null--cone is linked with the averaged Ricci scalar of space--like hypersurfaces according to (\\ref{eq:defkappa}). Unfortunately, the current available data are not sufficiently precise to unambiguously show a preferred selection of this hypothesis; this analysis should be done again with future data providing more statistics. On the other hand, we have calculated the explicit form of a function $C_\\CD (z_{D})$, previously introduced in \\cite{chris} to measure possible violations of the Copernican principle. This form can be considered as a {\\it prediction} of the particular models studied in this paper, and it will be a crucial test demonstrating a quantitatively significant difference to the standard FLRW paradigm. We have shown that the future EUCLID satellite project might be able to distinguish between a FLRW template geometry and the template geometry with evolving curvature in an averaged model by using joint measurements of a geometrical quantity ($d_{A}(z)$) and a kinematical property ($H(z)$). The effective metric that we employed has been motivated by physical and mathematical arguments, but it cannot be considered as a fully satisfactory description of the lightcone. In particular, since weak lensing involves a series of local effects of the metric on lightrays rather than just an integrated effect, implementing constraints from weak lensing surveys require a more refined study of the lightcone structure and of its link with spatial averaging. This link can be achieved by implementing the averaging formalism directly on the lightcone; this is the subject of work in progress. Another improvement will come from a closure condition that is better than a simple scaling solution, and that will encode more precisely the time evolution of backreaction. Such a closure condition can be looked for in numerical simulations, analytical approximations (see \\cite{marraetal1, marraetal2, rasanenpeaks} for particular approximations, and \\cite{wiltshire,wiltshire05,wiltshire07} for an interesting perspective), or observations (see \\cite{buchertcarfora:Q} and references therein for remarks on the difficulties of this last approach). Moreover, the complete study of other observables like the full CMB spectrum is still unavailable and will be crucial for the construction and the test of a `concordance model' for averaged inhomogeneous cosmologies. \\subsection*{\\sl Note added during revision} During the revision of this paper, \\cite{Rosenthal} published a preprint in which they pointed out that, in the effective model with a time varying curvature, the redshift can be calculated from first principles, as it is now done in Section 3.3 of this paper. This refinement of the redshift calculation indeed introduces quantitative changes, actually enhancing the differences between the model presented here and a FLRW model (as compared to the former version of or paper), but does not affect the conclusions of the paper. A few comments are in order to clarify the differences between our analysis and the one proposed in \\cite{Rosenthal}. One must note that instead of performing a full MCMC analysis of the model, the authors of \\cite{Rosenthal} fitted their model to a $\\Lambda$CDM fiducial model. Although it leads to correct models that actually reproduce the data (because $\\Lambda$CDM is a very good fitting model), it is by no way guaranteed that such a procedure provides all the acceptable values for the parameters, nor that it provides the most probable ones, as clearly shown by our analysis. \\subsection*{\\sl Acknowledgements} {\\small JL is supported by a Claude Leon Foundation Postdoctoral Fellowship. JL thanks C. Clarkson and B. Bassett for many valuable discussions and for having pointed out the use of $C$ as a test of the Copernican principle. TB acknowledges support and hospitality by Observatoire de Paris and Universit\\'e Paris 7, as well as by Universit\\'e de Gen\\`eve during working visits. MK acknowledges funding by the Swiss SNF for part of this work and thanks Syksy R\\\"as\\\"anen for valuable discussions.}" }, "0808/0808.3164_arXiv.txt": { "abstract": "The detection and quantification of narrow emission lines in X-ray spectra is a challenging statistical task. The Poisson nature of the photon counts leads to local random fluctuations in the observed spectrum that often results in excess emission in a narrow band of energy resembling a weak narrow line. From a formal statistical perspective, this leads to a (sometimes highly) multimodal likelihood. Many standard statistical procedures are based on (asymptotic) Gaussian approximations to the likelihood and simply cannot be used in such settings. Bayesian methods offer a more direct paradigm for accounting for such complicated likelihood functions but even here multimodal likelihoods pose significant computational challenges. The new Markov chain Monte Carlo (MCMC) methods developed in \\citet{vand:park:08} and \\citet{park:vand:08}, however, are able to fully explore the complex posterior distribution of the location of a narrow line, and thus provide valid statistical inference. Even with these computational tools, standard statistical quantities such as means and standard deviations cannot adequately summarize inference and standard testing procedures cannot be used to test for emission lines. In this paper, we use new efficient MCMC algorithms to fit the location of narrow emission lines, we develop new statistical strategies for summarizing highly multimodal distributions and quantifying valid statistical inference, and we extend the method of posterior predictive p-values proposed by \\citet{prot:etal:02} to test for the presence of narrow emission lines in X-ray spectra. We illustrate and validate our methods using simulation studies and apply them to the {\\it Chandra} observations of the high redshift quasar PG1634+706. ", "introduction": "\\label{park:sec:intro} \\subsection{Scientific Background} \\label{park:sec:sci} Modern X-ray observations show complex structures in both the spatial and spectral domains of various astrophysical sources. Nonetheless, active galalactic nuclei (AGN) including quasars' nuclei remain spatially unresolved even with the highest-resolution X-ray telescopes. Most of their energy is released within the unresolved core, and only spectral and timing information is available to study the nature of the X-ray emission. Generally speaking, emission and absorption lines constitute an important part of the X-ray spectrum in that they can provide information as to the state of plasma. One of the goals of X-ray data analysis is to understand the components present in the spectrum, and to obtain information about the emission and absorption features, as well as their locations and relation to the primary quasar emission. The detection of weak lines in noisy spectra is the main statistical problem in such analyses: Is a bump observed in the spectrum related to a real emission line or is it simply an artifact of the Poissonian noise? Although quasars' X-ray spectra are usually featureless as expected based on the Comptonization process \\citep[see for example][]{mark:nowa:wilm:05,sobo:siem:zyck:04,siko:etal:97}, an important X-ray emission feature identified in AGN and quasars spectra is the iron K emission line \\citep[see recent review by][]{miller:07}. Determining the origin and the nature of this line is one of main issues in AGN and quasar research. This line is thought to come directly from illuminated accretion flow as a fluorescent process \\citep{fabi:06}. The location of the line in the spectrum indicates the ionization state of iron in the emitting plasma, while the width of the line tells us the velocity of the plasma \\citep{fabi:06}. The iron line provides a direct probe of the innermost regions of accretion flow and matter in close vicinity of a black hole. Absorption features associated with the outflowing matter (warm wind, partial covering absorber) have also been observed in recent X-ray observations \\citep{gall:etal:02,char:etal:02,poun:reev:07}. Although the location and width of absorption lines provide information as to the velocity of the absorber and its distance from the quasar, this article focuses on statistical issues in fitting the spectral location of narrow emission lines, i.e., identifying the ionization state. There are two parts to the Fe-K-alpha emission line observed in AGN \\citep{yaqoob:etal:01}: one is a broad component thought to be a signature of a relativistic motion in the innermost regions of an accretion flow; the other is a narrow component that is a result of a reflection off the material at larger distances from the central black hole. A detection of the broad component is challenging as it requires a spectral coverage over a large energy range, so the continuum emission is well determined and the broad line can be separated \\citep{reeves:etal:06}. The relativistic line profile is broad and skewed, and two strong peaks of the emission line that originates in a relativistic disk can be prominent and narrow. While the full profile of the broad line may not be easily separable from the continuum, these two peaks may provide a signature for this line in the X-ray spectrum. The broad Fe-line gives an important diagnostic of the gas motion and can be used to determine the spin of a black hole \\citep{miller:06}; see also an alternative model for the ``red wing\" component by \\citet{miller:etal:08}. The narrow component of the line gives diagnostics of the matter outside the accretion disk and conditions at larger distances from the black hole; see Fe-line Baldwin effect discussion in \\citet{jiang:etal:06}. Both line components are variable and the line may ``disappear'' from the spectrum \\citep{yaq:etal:01}. The spectral resolution of X-ray CCD detectors (for example 100-200~eV in ACIS on {\\it Chandra} or EPIC on XMM-{\\it Newton}) is relatively low with respect to the predicted width of narrow ($< 5000-30,000$~km~s$^{-1}$) emission or absorption lines in AGN and quasars. Observations with grating instruments (RGS or HEG) can provide high resolution X-ray spectra, but the effective area of the present X-ray telescopes is too low for efficient AGN detections, and only a handful of bright low redshift sources have been observed with gratings to date \\citep{yaqoob:07}. Therefore mainly the X-ray CCD spectra of lower resolution are used to study large samples of AGN and quasars \\citep[see for example][]{guai:06,page:etal:05,jimen:etal:05}. Using these relatively low resolution X-ray detectors, the Fe-K-alpha emission line can be narrow enough to be contained entirely in a single detector bin. In some cases (for example in {\\it Chandra}) the line may occupy a few bins. In this article we focus on the statistical problem of fitting the spectral location of an emission line or a set of emission lines that are narrow. This is a common objective in high-energy analyses, but as we shall discuss fitting these relatively narrow features poses significant statistical challenges. In particular we find evidence that using line profiles that are narrower than we actually expect the emission line to be can improve the statistical properties of the fitted emission line location. \\subsection{X-Ray Spectral Analysis} \\label{park:sec:hea} X-ray spectra, such as those available with the {\\it Chandra X-ray Observatory} carry much information as to the quasar's physics. Taking advantage of the spectral capacity of such instruments, however, requires careful statistical analysis. For example, the resolution of such instruments corresponds to a fine discretization of the energy spectrum. As a result, we expect a low number of counts in each bin of the X-ray spectrum. Such low-count data make the Gaussian assumptions that are inherent in traditional minimum $\\chi^2$ fitting inappropriate. A better strategy, which we employ, explicitly models photon arrivals as an inhomogeneous Poisson process \\citep{vand:etal:01}. In addition, data are subject to a number of processes that significantly degrade the source counts, e.g., the absorption, non-constant effective area, blurring of photons' energy, background contamination, and photon pile-up. Thus, we employ statistical models that directly account for these aspects of data collection. In particular, we design a highly structured multilevel spectral model with components for both the data collection processes and the complex spectral structures of the sources themselves. In this highly structured spectral model, a Bayesian perspective renders straightforward methods that can handle the complexity of {\\it Chandra} data \\citep{vand:etal:01,vand:kang:04,vand:etal:06}. As we shall illustrate, these methods allow us to use low-count data, to search for the location of a narrow spectral line, to investigate its location's uncertainty, and to construct statistical tests that measure the evidence in the data for including the spectral line in the source model. \\subsection{A Statistical Model for the Spectrum} \\label{park:sec:model} The energy spectrum can be separated into two basic parts: a set of continuum terms and a set of several emission lines\\footnote{The model can be extended to account for absorption lines, but in this paper we focus on additive features such as emission lines.}. We begin with a standard spectral model that accounts for a single continuum term along with several spectral lines. Throughout this paper, we use $\\theta$ as a general representation of model parameters in the spectral model. The components of $\\theta=(\\theta^C,\\theta^L,\\theta^A,\\theta^B)$ represent the collection of parameters for the Continuum, (emission) Lines, Absorption, and Background contamination, respectively. (Notice that the roman letters in the superscripts serve as a mnemonic for these four processes.) Because the X-ray emission is measured by counting the arriving photons, we model the expected Poisson counts in energy bin $j\\in{\\cal J}$, where ${\\cal J}$ is the set of energy bins, as \\begin{eqnarray} \\Lambda_j(\\theta) \\ =\\ \\Delta_j f\\big(\\theta^C,E_j\\big)+ \\sum\\limits_{k=1}^K\\lambda_k\\pi_j\\big(\\mu_k,\\nu_k\\big), \\label{park:eq:ideal} \\end{eqnarray} where $\\Delta_j$ and $E_j$ are the width and mean energy of bin $j$, $f(\\theta^C,E_j)$ is the expected counts per unit energy due to the continuum term at energy $E_j$, $\\theta^C$ is the set of free parameters in the continuum model, $K$ is the number of emission lines, $\\lambda_k$ is the expected counts due to the emission line $k$, and $\\pi_j(\\mu_k,\\nu_k)$ is the proportion of an emission line centered at energy $\\mu_k$ and with width $\\nu_k$ that falls into bin $j$. There are a number of smooth parametric forms to describe the continuum in some bounded energy range; in this article we parameterize the continuum term $f$ as a power law, i.e., $f(\\theta^C,E_j)=\\alpha^CE_j^{-\\beta^C}$ where $\\alpha^C$ and $\\beta^C$ represent the normalization and photon index, respectively. The emission lines can be modeled via the proportions $\\pi_j(\\mu_k,\\nu_k)$ using narrow Gaussian distributions, Lorentzian distributions, or delta functions; the counts due to the emission line are distributed among the bins according to these proportions. While the Gaussian or Lorentzian function parameterizes an emission line in terms of center and width, the center is the only free parameter with a delta function; the width of the delta function is effectively the width of the energy bin in which it resides. While the model in Equation~\\ref{park:eq:ideal} is of primary scientific interest, a more complex statistical model is needed to address the data collection processes mentioned in \\S\\ref{park:sec:hea}. We use the term {\\it statistical model} to refer to the model that combines the {\\it source} or {\\it astrophysical model} with a model for the stochastic processes involved in data collection and recording. Thus, in addition to the source model, the statistical model describes such processes as instrument response and background contamination. Specifically, to account for the data collection processes, Equation~\\ref{park:eq:ideal} is modified via \\begin{eqnarray} \\Xi_l(\\theta) \\ =\\ \\sum\\limits_{j\\in{\\cal J}} M_{lj}\\Lambda_j(\\theta)d_j u(\\theta^A,E_j)+\\theta_l^B \\label{park:eq:obs} \\end{eqnarray} where $\\Xi_l(\\theta)$ is the expected observed Poisson counts in detector channel $l\\in{\\cal L}$, ${\\cal L}$ is the set of detector channels, $M_{lj}$ is the probability that a photon that arrives with energy corresponding to bin $j$ is recorded in detector channel $l$ (i.e., $\\mathbf{M}=\\{M_{lj}\\}$ is the so-called redistribution matrix or RMF commonly used in X-ray analysis), $d_j$ is the effective area (i.e., ARF, a calibration file associated with the X-ray observation) of bin $j$, $u(\\theta^A,E_j)$ is the probability that a photon with energy $E_j$ is {\\it not} absorbed, $\\theta^A$ is the collection of parameters for absorption, and $\\theta_l^B$ is a Poisson intensity of the background counts in channel $l$. While the scatter probability $M_{lj}$ and the effective area $d_j$ are presumed known from calibration, the absorption probability is parameterized using a smooth function; see \\citet{vand:hans:02} for details. To quantify background contamination, a second data set is collected that is assumed to consist only of background counts; the background photon arrivals are also modeled as an inhomogeneous Poisson process. \\subsection{Difficulty with Identifying Narrow Emission Lines} \\label{park:sec:narrow} Unfortunately, the statistical methods and algorithms developed in \\citet{vand:etal:01} cannot be directly applied to fitting {\\it narrow} emission lines. There are three obstacles that must be overcome in order to extend Bayesian highly structured models to spectra containing narrow lines. In particular, we must develop (1) new computational algorithms, (2) statistical summaries and methods for inference under highly multimodal posterior distributions, and (3) statistical tests that allow us to quantify the statistical support in the data for including an emission line or lines in the model. Our main objective in this paper is to extend the methods of \\citet{vand:etal:01} in these three directions, and to evaluate and illustrate our proposals. Here we discuss each of these challenges in detail. \\paragraph{Challenge 1: Statistical Computation.} Fitting the location of narrow lines requires new and more sophisticated computational techniques than those developed by \\citet{vand:etal:01}. Indeed, the algorithms that we develop require a new theoretical framework for statistical computation: they are not examples of any existing algorithm with known properties. Although the details of this generalization are well beyond the scope of this article, we can offer a heuristic description; a more detailed description is given in Appendix~\\ref{ap:alg}. Readers who are interested in the necessary theoretical development of the statistical computation techniques are directed to \\citet{vand:park:04,vand:park:08} and \\citet{park:vand:08}. The algorithms used by \\citet{vand:etal:01} to fit the structured Bayesian model described in \\S\\ref{park:sec:model} are based on the probabilistic properties of the statistical models. For example, the parameters of a Gaussian line profile can be fit by iteratively attributing a subset of the observed photons to the line profile and using the mean and variance of these photon energies to update the center and width of the line profile. The updated parameters of the line profile are used to again attribute a subset of the photons to the line, i.e., to stochastically select a subset of the photons that are likely to have arisen out of the physical processes at the source corresponding to the emission line. These algorithms are typically very stable. For example, they only return statistically meaningful parameters because the algorithms themselves mimic the probabilistic characteristics of the statistical model. The family of Expectation/Maximization (EM) algorithms \\citep{demp:lair:rubi:77} and Markov chain Monte Carlo (MCMC) methods such as the Gibbs sampler \\citep{gema:gema:84} are the examples of statistical algorithms of this sort. A drawback of these algorithms is that in some situations they can be slow to converge. When fitting the location of a Gaussian emission line, for example, the location is updated more slowly if the line profile is narrower. This is because only photons with energies very close to the current value of the line location can be attributed to the line. Updating the line location with the mean of the energies of these photons cannot result in a large change in the emission line location. The situation becomes chronic when a delta function is used to model the line profile: The line location parameter sticks at its starting value throughout the iteration. It is to circumvent this difficulty that we develop both new EM-type algorithms \\citep{vand:park:04} and new MCMC samplers specially tailored for fitting narrow lines. Our new samplers are motivated by the Gibbs sampler, but constitute a non-trivial generalization of Gibbs sampling known as partially collapsed Gibbs sampling \\citep{vand:park:08,park:vand:08}; see Appendix~\\ref{ap:alg}. Our updated versions of both classes of algorithms are able to fit narrow lines by avoiding the attribution of photons to the emission line during the iteration. Such algorithms tend to require fewer iterations to converge regardless of the width of the emission line. Because they involve additional evaluation of quantities evolving the large dimensional redistribution matrix, $M$, however, each iteration of these algorithms can be significantly more costly in terms of computing time. A full investigation of the relative merit of the algorithms and a description of how the computational trade-offs can be played to derive optimal algorithms are beyond the scope of this paper. Except in Appendix~\\ref{ap:alg}, we do not discuss the details of the algorithms further in this article; interested readers are directed to \\citet{vand:park:04,vand:park:08} and \\citet{park:vand:08}. \\paragraph{Challenge 2: Multimodal Likelihoods.} The likelihood function for the emission line location(s) is highly multimodal. Each mode corresponds to a different relatively likely location for an emission line or a set of emission lines. Standard statistical techniques such as computing the estimates of the line locations with their associated error bars or confidence intervals implicitly assume that the likelihood function is unimodal and bell shaped. Because this assumption is clearly and dramatically violated, these standard summary statistics are unreliable and inadequate. Unfortunately, there are no readily available and generally applicable simple statistical summaries to handle highly multimodal likelihoods. Instead we must develop summaries that are tailored to the specific scientific goals in a given analysis. Because general strategies for dealing with multimodal likelihood functions are little known to astronomers and specific strategies for dealing with multimodal likelihood functions for the location of narrow spectral lines do not exist, one of the primary goals of this article is to develop and illustrate these methods. A fully Bayesian analysis of our spectral model with narrow emission lines is computationally demanding, even with our new algorithms. Thus, we develop techniques that are much quicker and give similar results for the location of emission lines. These methods based on the so-called {\\it profile posterior distribution} do not stand on as firm of a theoretical footing as a fully Bayesian analysis, but are much quicker and thus better suited for {\\it exploratory data analysis}. The profile posterior distribution along with our exploratory methods are fully described and compared with the more sophisticated Bayesian analysis. \\paragraph{Challenge 3: Testing for the Presence of Narrow Lines.} In addition to fitting the location of one or more emission lines, we often would like to perform a formal test for the inclusion of the emission lines in the statistical model. That is, we would like to quantify the evidence in a potentially sparse data set for a particular emission line in the source. Testing for a spectral line is an example of a notoriously difficult statistical problem in which the standard theory does not apply. There are two basic technical problems. First, the simpler model that does not include a particular emission line is on the boundary of the larger model that does include the line. That is, the intensity parameter of an emission line is zero under the simpler model and cannot be negative under the larger model. An even more fundamental problem occurs if either the line location or width is fit, because these parameters have {\\it no value} under the simpler model. The behavior (i.e., sampling distribution) of the likelihood ratio test statistic under the simpler model is not well understood and cannot be assumed to follow the standard $\\chi^2$ distribution, even asymptotically. \\citet{prot:etal:02} propose a Monte-Carlo-based solution to this problem based on the method of posterior predictive p-values \\citep{rubi:84,meng:94a}. In this article we extend the application of Protassov~{\\it et~al}.'s solution to the case when we fit the location of a narrow emission line, a situation that was avoided in \\citet{prot:etal:02}. \\subsection{Outline of the Article} \\label{park:sec:outline} The remainder of the article is organized into four sections. \\S\\ref{park:sec:model-based} reviews Bayesian inference and Monte Carlo methods with an emphasis on multimodal distributions, outlines our computation methods, proposes new summaries of multimodal distributions, and describes exploratory statistical methods in this setting. We introduce illustrative examples in \\S\\ref{park:sec:model-based}, but detailed spectral analysis is postponed in order to allow us to focus on our proposed methods. In \\S\\ref{park:sec:simul}, a simulation study is performed to investigate the statistical properties of our proposed methods, with some emphasis placed on the potential benefits of model misspecification. \\S\\ref{park:sec:quasar} presents the analysis of the high redshift quasar PG1634+706, and how to test for the inclusion of the line in the spectral model. Concluding remarks appear in \\S\\ref{park:sec:conclusion}. An appendix outlines the computational methods we developed specifically for fitting the location of narrow emission lines. ", "conclusions": "\\label{park:sec:conclusion} This article presents methods to detect, identify, and locate narrow emission lines in X-ray spectra via a highly structured multilevel spectral model that includes a delta function line profile. Modeling narrow emission lines with a delta function causes the EM algorithms and MCMC samplers developed in \\citet{vand:etal:01} to break down and thus requires more sophisticated statistical methods and algorithms. The marginal posterior distribution of the delta function emission line location tends to be highly multimodal when the emission line is weak or multiple emission lines are present in the spectrum. Because basic summary statistics are not appropriate to summarize such a multimodal distribution, we instead develop and use HPD graphs along with a list of posterior modes. Testing for an emission line in the spectrum is a notoriously challenging problem because the value of the line intensity parameter is on the boundary of the parameter space, i.e., zero, under a model that does not include an emission line. Thus, we extend the posterior predictive methods proposed by \\citet{prot:etal:02} to test for the evidence of a delta function emission line with unknown location in the spectrum. Using the simulation study in \\S\\ref{park:sec:simul}, we demonstrate the potential advantage of model misspecification using a delta function line profile in place of a Gaussian line profile. We show that the delta function line profile may provide more precise and meaningful summaries for line locations if the true emission line is narrow. When multiple lines are present in the spectrum, the marginal posterior distribution of a single delta function line location may indicate multiple lines in the spectrum. Our methods are applied to the six different {\\it Chandra} observations of PG1634+706 in order to identify a narrow emission line in the X-ray spectrum. Given all the six observations, the most probable delta function line is identified at $2.845_{-0.055}^{+0.045}$~keV in the observed frame. The corresponding rest frame energy for the line is $6.64_{-0.13}^{+0.11}$~keV, which may suggest the high ionization of iron in the emitting plasma. There is some recent evidence that high ionization iron line can be variable on short timescales \\citep[see for example Mkn766 in][]{miller:etal:06}. Such variability would explain no detection of the emission line in one of the six {\\it Chandra} observations." }, "0808/0808.3022_arXiv.txt": { "abstract": "We observed an area of 10 deg$^2$ of the Large Magellanic Cloud using the Infrared Camera on board AKARI. The observations were carried out using five imaging filters (3, 7, 11, 15, and 24 micron) and a dispersion prism (2 -- 5 micron, $\\lambda / \\Delta\\lambda$ $\\sim$ 20) equipped in the IRC. This paper describes the outline of our survey project and presents some initial results using the imaging data that detected over 5.9$\\times$10$^5$ near-infrared and 6.4$\\times$10$^4$ mid-infrared point sources. The 10 $\\sigma$ detection limits of our survey are about 16.5, 14.0, 12.3, 10.8, and 9.2 in Vega-magnitude at 3, 7, 11, 15, and 24 micron, respectively. The 11 and 15 micron data, which are unique to AKARI IRC, allow us to construct color-magnitude diagrams that are useful to identify stars with circumstellar dust. We found a new sequence in the color-magnitude diagram, which is attributed to red giants with luminosity fainter than that of the tip of the first red giant branch. We suggest that this sequence is likely to be related to the broad emission feature of aluminium oxide at 11.5 micron. The 11 and 15 micron data also indicate that the ([11] $-$ [15]) micron color of both oxygen-rich and carbon-rich red giants once becomes blue and then turns red again in the course of their evolution, probably due to the change in the flux ratio of the silicate or silicon carbide emission feature at 10 or 11.3 micron to the 15 micron flux. ", "introduction": "Owing to its proximity ($\\sim$ 50 kpc; e.g., \\cite{feast1987}) and face-on geometry, the Large Magellanic Cloud (LMC) is an ideal natural laboratory for the study of various astrophysical fields. It is located at a high Galactic latitude of $\\sim$ $-$36$^\\circ$, and we expect less contamination of foreground stars and less interstellar extinction. It is far enough to neglect its depth so that we can reasonably assume that objects in the LMC are all at the same distance from us. The apparent size of the LMC is also in a size, for which the entire galaxy can be surveyed in a reasonable amount of time to study the material circulation processes and star-formation history in a galactic scale. In the meanwhile, it is close enough to resolve and study individual objects even with relatively small telescopes. Moreover, the mean metallicity of the LMC is known to be small ($\\sim$ 1/4) compared to the solar, intriguing us to study the influence of low metallicity on various astrophysical phenomena. \\begin{figure}[htbp] \\begin{center} \\FigureFile(86mm,86mm){figure1bw.ps} % \\end{center} \\caption{Observed area of the AKARI IRC survey (thick solid outline), overlaid on the photographic image kindly provided by Mr. Motonori Kamiya. The thick dashed outline indicates the coverage of the \\textit{Spitzer}/SAGE survey (7$^\\circ$$\\times$7$^\\circ$, \\cite{meixner2006}), the thick dash-doted outline shows the coverage of the IRSF/SIRIUS near-infrared survey (6.3$^\\circ$$\\times$6.3$^\\circ$, \\cite{kato2007}), and the thin solid outline represents the coverage of the Magellanic clouds optical photometric survey (8.5$^\\circ$$\\times$7.5$^\\circ$, \\cite{zaritsky2004}).} \\label{fig:surveyregion} \\end{figure} Because of these characteristics there have been a number of survey projects of the LMC in various wavelengths (Hard X-ray: \\cite{gotz2006}; Soft X-ray: e.g., \\cite{long1981} with Einstein, \\cite{sasaki2000}, \\cite{haberl1999} with ROSAT, and \\cite{haberl2003} with XMM; UV: \\cite{smith1987}; H$\\alpha$: \\cite{kennicut1986}, \\cite{gaustad2001}; Optical: \\cite{zaritsky2004}; NIR: e.g., \\cite{cioni2000a} with DENIS, \\cite{nikolaev2000} with 2MASS, and \\cite{kato2007} with IRSF/SIRIUS; MIR\\&FIR: e.g., \\cite{israel1986} with IRAS, \\cite{egan2003} with MSX, and \\cite{meixner2006} with \\textit{Spitzer}/SAGE; [CII]: \\cite{mochizuki}; CO: e.g., \\cite{mizuno2001} with NANTEN; HI: \\cite{luks1992}, \\cite{kim1998}). A ground-based near-infrared (NIR) survey detected about fifteen million sources in the $\\sim$ 40 deg$^2$ area of the LMC (\\cite{kato2007}). On the other hand, previous mid-infrared (MIR) surveys detected only a few thousands of sources in $\\sim$ 100 deg$^2$ area of the LMC (\\cite{egan2003}), which is too shallow to compare with the survey observations in other wavelengths. Moreover, the angular resolutions of previous MIR surveys were not high enough to make secure cross-identifications with other survey catalogs. This situation is significantly improved with the advent of the \\textit{Spitzer Space Telescope} (SST; \\cite{werner2004}) and AKARI, which have instruments capable of deep mid-infrared observations with the high spatial resolution. The \\textit{Spitzer} SAGE project (\\cite{meixner2006}) carried out an uniform and unbiased imaging survey of about 49 deg$^2$ area of the LMC, providing photometry for about 2$\\times$10$^5$ sources with all IRAC \\citep{fazio2004a} bands at two epochs separated by 3 months. We carried out near- to mid-infrared imaging and near-infrared spectroscopic survey toward the LMC with AKARI. The entire LMC was also observed as part of the All-Sky survey at 6 bands in the mid- to far-infrared with AKARI (\\cite{ishihara2006}; \\cite{kawada2007}). AKARI observations provide multi-band (11 bands) data of the LMC from the near- to far-infrared (FIR), which will give us a new insight into the various phenomena occurring in the LMC. In this paper, we provide an overview of our LMC survey project and present some initial results using the preliminary photometric catalog of bright point sources. Then we will make general analyses using the catalog with an emphasis on the unique characteristics of the IRC observations compared to {\\it Spitzer} observations. The first AKARI/IRC LMC point source catalog is planned to be released to the public in 2009. ", "conclusions": "We carried out an imaging and spectroscopic survey of a 10 deg$^2$ area of the LMC using IRC onboard AKARI. In this paper we discuss the imaging data obtained in the survey and describe a preliminary photometric catalog of bright point sources. The first point source catalog is planned to be released to the public in 2009. We cross-identified our catalog with the existing optical, near-infrared, and mid-infrared photometry catalogs, and used the combined data to build color-magnitude diagrams using several combinations of interesting wavebands and also to calculate absolute bolometric magnitudes. By virtue of the S11 and L15 bands that are unique to IRC, we found new interesting features in the color-magnitude diagrams. The present IRC data and the upcoming results from AKARI All-Sky survey together with existing radio and near-infrared ground-based observational results will provide a significant database to study the star-formation history and material circulation within a galaxy." }, "0808/0808.3813_arXiv.txt": { "abstract": "We present the first results of a multi-object spectroscopic (MOS) campaign to follow up cluster candidates located via weak lensing. Our main goals are to search for spatial concentrations of galaxies that are plausible optical counterparts of the weak lensing signals, and to determine the cluster redshifts from those of member galaxies. Around each of 36 targeted cluster candidates, we obtain $15-32$ galaxy redshifts. For 28 of these targets, we confirm a secure cluster identification, with more than five spectroscopic galaxies within a velocity of $\\pm 3000$km/s. This includes three cases where two clusters at different redshifts are projected along the same line-of-sight. In 6 of the 8 unconfirmed targets, we find multiple small galaxy concentrations at different redshifts, each containing at least three spectroscopic galaxies. The weak lensing signal around those systems is thus probably created by the projection of groups or small clusters along the same line-of-sight. In both the remaining two targets, a single small galaxy concentration is found. We evaluate the weak lensing mass of confirmed clusters via two methods: aperture densitometry and by fitting to an NFW model. In most cases, these two mass estimates agree well. In some candidate super-cluster systems, we find additional evidence of filaments connecting the main density peak to additional nearby structure. For a subsample of our most cleanly measured clusters, we investigate the statistical relation between their weak lensing mass ($M_{\\rm NFW}$, $\\sigma_{\\rm sis}$) and the velocity dispersion of their member galaxies ($\\sigma_{\\rm v}$), comparing our sample with optically and X-ray selected samples from the literature. Our lensing-selected clusters are consistent with $\\sigma_v = \\sigma_{\\rm sis}$, with a similar scatter to the optically and X-ray selected clusters. We thus find no evidence of selection bias compared to these other techniques. We also derive an empirical relation between the cluster mass and the galaxy velocity dispersion, $M_{200}=9.6\\times 10^{14}\\times (\\sigma_v/1000$km/s$)^{2.7}/E(z) h^{-1}M_\\odot$, which is in reasonable agreement with the prediction of $N$-body simulations in the $\\Lambda$CDM cosmology. ", "introduction": "\\label{sec:intro} The development of weak lensing techniques, coupled with deep panoramic imaging surveys, has enabled us to locate clusters of galaxies via the gravitational distortion of background galaxies' shapes. Since the first, spectroscopically confirmed discovery of a shear-selected cluster by Wittman et al. (2001), there has been rapid progress toward a large, weak-lensing selected cluster catalogue. Miyazaki et al. (2003) first reported the detection of several significant shear-selected cluster candidates in an untargeted 2.1 deg$^2$ field. Hetterscheidt et al. (2005) found 5 cluster candidates in 50 randomly selected VLT FORS1 fields (0.64 deg$^2$ in total), all of which are associated with an overdensity of galaxies. Wittman et al. (2006) reported 8 candidates in the 8.6 deg$^2$ Deep Lens Survey. Gavazzi \\& Soucail (2007) found 14 cluster candidates in the 4 deg$^2$ CFHT Legacy Survey (Deep), of which nine have optical or X-ray counterparts and are thus secure clusters. The first sizable sample of weak lensing shear-selected cluster candidates was presented by Miyazaki et al. (2007; hereafter P1). Their sample was obtained solely via peak finding in weak lensing density maps, and includes 100 significant peaks in a 16.7 deg$^2$ survey area. Before such a sample is used for statistical cosmological or astronomical analyses, two additional follow-up observations are required. Firstly, each cluster candidate should be confirmed by independent observations, since a fraction of lensing peaks could be false positives from e.g.\\ the chance tangential alignment of galaxies' intrinsic ellipticities (White, van Waerbeke \\& Mackey 2002; Hamana, Takada \\& Yoshida, 2004; Hennawi \\& Spergel 2005). Secondly, the redshifts of confirmed clusters need to be determined in order to derive their physical quantities, including mass. We have conducted a multi-object spectroscopic (MOS) campaign that accomplishes both goals. We have measured the redshifts of a few tens of galaxies within an expected cluster scale radius (or core radius, typically a few arcmins), and searched for spatial concentrations that are plausible optical counterparts of the weak lensing signals. Once a galaxy overdensity is found, it is easy to determine the cluster redshift from the redshifts of member galaxies. It is important to note that cluster confirmation based on prominent galaxy concentrations would not be very effective for very high mass-to-light ratio ($M/L$), galaxy-poor systems. Although our methodology can {\\it confirm} normal or galaxy-rich clusters, the absence of a galaxy concentration in our fairly sparsely-sampled data therefore does not necessarily prove that a weak lensing signal is false. In addition to our primary goals, multi-object spectroscopic observations provide several useful by-products. If redshifts can be obtained for sufficient galaxies in a cluster, their line-of-sight velocity dispersion provides an estimate of the cluster's dynamical mass. MOS observations can also detect multiple structures along the same line of sight. Because of the relatively broad redshift window function of gravitational lensing, physically unrelated structures in the same line of sight may contribute to a single peak in a weak lensing density map, resulting in an overestimation of the cluster mass (White et al. 2002). It will therefore be important to quantify and properly account for such projections when computing statistics of cluster masses from weak lensing observations. In this paper, we present results of cluster confirmations and cluster redshifts. We discuss the detailed weak lensing properties of each system and, for a clean subset of our clusters, examine statistical relations between the weak lensing masses and dynamical masses. A statistical analysis of the {\\it entire} sample, taking into account additional selection effects, will be presented in Green et al. (in preparation). This paper is organized as follows. In \\S2, we discuss our selection of cluster candidates. In \\S3, we describe our new observations, data reduction and measurements of galaxy redshifts. In \\S4, we identify optical counterparts to cluster candidates, and measure their velocity dispersions and dynamical masses. In \\S5, we analyze the weak lensing signal of confirmed clusters. In \\S6, we investigate cluster scaling relations within our sample. In \\S7, we summarize our results. In Appendix~\\ref{appendix:nfw}, we calculate the gravitational lensing shear profile of a truncated NFW model. Detailed discussions of each system, including comparisons of the dynamic and lensing masses, follow in Appendices~\\ref{appendix:prop} (for clusters we have observed) and \\ref{appendix:xmmlss} (for observations taken from the literature). Throughout this paper, we adopt a flat $\\Lambda$CDM cosmology with the matter density $\\Omega_{\\rm m}=0.3$, the cosmological constant $\\Omega_\\Lambda=0.7$, the Hubble constant $H_0=100 h$ km~s$^{-1}$~Mpc$^{-1}$ with $h=0.7$. \\begin{table*} \\caption{Summary of spectroscopic observations. (a) Cluster name in the IAU convention. (b) Field and catalogue number given in P1 (if listed). (c) The peak $\\kappa$ value. (d) The number density of galaxies with $18$20 keV; e.g. \\citealt{Sguera2006}) or at softer energy bands (\\citealt{Smith2006aa}, \\citealt{zand2005}). The only SFXT previously observed simultaneously in a wide X--ray band was IGR~J16479--4514, during a flare caught with the {\\it Swift} satellite \\citep{Romano2008:sfxts_paperII}. The X--ray spectroscopy shows that these two SFXTs, which are considered the prototypes of this new class of HMXBs, have different properties during the bright flares. \\src\\ is one order of magnitude less absorbed than \\srcxte, and displays a significantly flatter spectrum below 10 keV, with a XRT/WT spectrum well fitted with a power-law with a photon index of 0.75$\\pm{0.11}$, compared with the \\srcxte\\ photon index, which lies in the range 1--2. The 1--10 keV spectral properties observed in \\src\\ during the flare are similar to what observed previously with $Chandra$ \\citep{zand2005}, where the absorbed powerlaw fit resulted in a photon index of 0.73$\\pm{0.13}$, a column density of (1.36$\\pm{0.22}$)$\\times$10$^{22}$~cm$^{-2}$, and a peak flux of $\\sim$3$\\times$10$^{-9}$~erg~cm$^{-2}$~s$^{-1}$. The broad band analysis shows that \\src\\ displays a quite sharp cutoff at 18$\\pm{2}$~keV (when using the power law model with a high energy cut-off, {\\sc highecut} in Table~\\ref{tab:specigr}) or a well constrained temperature for the Comptonizing electrons (in the {\\sc comptt} model in XSPEC) at 4--5 keV. Instead, in \\srcxte, a single power law (photon index of 2.2--2.5) can describe the whole spectrum from soft to hard energies. Part of this difference could be explained by the much higher absorption towards the line of sight of \\srcxte, which does not allow to constrain well the low energy part of the power-law model. \\begin{figure}[th!]% \\centerline{\\includegraphics[width=8.5cm,height=12cm,angle=0]{figure6.ps}} \\caption{Light curves of the outbursts of SFXTs followed by {\\it Swift}/XRT referred to their respective triggers. We show the 2005 outburst of IGR~J16479$-$4514 \\citep{Sidoli2008:sfxts_paperI}, which has a better coverage than the one observed in 2008 \\citep{Romano2008:sfxts_paperII}. The IGR~J11215$-$5952 light curve has an arbitrary start time, since the source did not trigger the BAT (the observations were obtained as a ToO; \\citet{Romano2007}. Note that where no data are plotted, no data were collected. For clarity, the time interval between two consecutive dashed vertical lines is one day. } [See the electronic edition of the Journal for a color version of this figure.] \\label{fig:comp} \\end{figure} The observations we are reporting here are part of an on-going monitoring campaign of four SFXTs with {\\it Swift} \\citep{Sidoli2008:sfxts_paperI}, which started on 2007 October 26. The two bright flares discussed here are the first from these two SFXTs, since the start of the campaign, which could be simultaneously covered with both {\\it Swift} XRT and BAT. The results on the out-of-outburst X--ray emission (below 10 keV) have been reported by \\citet{Sidoli2008:sfxts_paperI}, where we find evidence that the accretion is still present, over long timescales of months, even outside the bright outbursts. Both \\srcxte\\ and \\src\\ show evidence that they still accrete matter even outside the outbursts, at a much fainter (100--1000 times lower) level than during the flares, with still a large flux variability (at least one order of magnitude). A complete view of the different luminosity and spectral states of the monitored SFXTs will be clearer at the end of the campaign, but it is already possible to compare the average out-of-outburst emission properties with the spectra during the flares. Regarding the 0.3--10 keV spectra (fitted with a simple absorbed power-law), \\srcxte\\ appears to be much more absorbed during the flare than during the out-of-outbust emission (see Fig.~\\ref{fig:contxte}), while the photon index is similar, within the large uncertainties \\citep{Sidoli2008:sfxts_paperI}. Similar changes in the absorbing column density of \\srcxte\\ have been observed before with $RXTE$/PCA and ASCA \\citep{Smith2006aa}, but during bright outbursts, where the $N_{\\rm H}$ ranged, from one bright flare to another, from 3 to 37$\\times10^{22}$~cm$^{-2}$. A $Chandra$ observation \\citep{Smith2006aa} displaying an unabsorbed 1--10~keV flux of $\\sim10^{-11}$~erg~cm$^{-2}$~s$^{-1}$, intermediate between the average out-of-outburst emission \\citep{Sidoli2008:sfxts_paperI} and the bright flare observed here, shows a hard power law spectrum with a photon index of 0.62$\\pm{0.23}$, absorbed with a column density $N_{\\rm H}$=(4.2$\\pm{1.0}$)$\\times10^{22}$~cm$^{-2}$, which is compatible with that of the out-of-outburst emission. Thus, in \\srcxte, there does not seem to be a clear correlation between source intensity, spectral hardness and absorbing column density, to date. Instead, \\src\\ shows a significantly harder spectrum during the flare, and a lower column density than during the out-of-outburst phase reported in \\citet{Sidoli2008:sfxts_paperI}, obtained summing together all the XRT data available from 2007 October 27 to 2008 February 28. During the out-of-outburst phase, the average observed flux was $\\sim$3$\\times10^{-12}$~erg~cm$^{-2}$~s$^{-1}$ (2--10 keV), the powerlaw photon index, $\\Gamma$, was 2.1$ ^{+0.6} _{-0.5}$, and the absorbing column density N$_{\\rm H}$=(3.2 $^{+1.2} _{-0.9}$)$\\times10^{22}$~cm$^{-2}$ \\citep{Sidoli2008:sfxts_paperI}. We also compared the bright flare spectroscopy with the spectrum extracted from one of the observations obtained a few days before the bright flare from \\src\\ (dashed contours in Fig.~\\ref{fig:cont}). A hardening of the \\src\\ spectrum during the flaring activity is evident. A similar behaviour was already suggested from the analysis of different XMM-Newton observations during a low level flaring activity \\citep{Gonzalez2004}, and from the analysis of the 2004 $Chandra$ observation \\citep{zand2005}. A proper comparison with the INTEGRAL results of a few outbursts from \\src\\ reported by Sguera et al. (2006) cannot be done since the energy range with INTEGRAL was limited to 20--60~keV. These authors fitted the 20--60~keV spectrum with a thermal bremsstrahlung, which is clearly not adequate to describe our XRT+BAT spectrum (reduced $\\chi^{2}_{\\nu}$ of 1.7, for 159 dof). Different mechanisms have been proposed to explain the bright and short duration flaring activity in this new class of sources. Some models are related to the structure of the supergiant companion wind, involving spherically simmetric clumpy winds (see e.g. \\citealt{zand2005}, \\citealt{Negueruela2008}) or anisotropic winds \\citep{Sidoli2007}; other models involve the interaction of the inflowing wind with the neutron star magnetosphere (see e.g. Bozzo et al. 2008). Sidoli et al. (2007) explain the outbursts as being due to enhanced accretion onto the neutron star when it crosses, moving along the orbit, an equatorial wind disk component from the supergiant companion. Depending on the thickness and truncation of this supposed disk wind and on its inclination with respect to the orbital plane of the binary system, the compact object will cross once or twice in a periodic or quasi-periodic manner the disk, undergoing outbursts. In the framework of this model, the geometry, the structure of this disk wind and its inclination with respect to the line of sight could explain the variability in the local absorbing column density, even during different outbursts (as observed several times in \\srcxte) and compared with the low level activity. A lower column density during the out-of-outburst activity could be due to the fact the source is completely outside the denser equatorial wind from the companion. In the spherically symmetric clumpy winds model, the difference in the observed column density could be due to the accreting dense clumps. On the other hand, in this case the clump matter should remain neutral also in proximity of the neutron star during bright flares. We think it is more likely that the absorbing column density is not related with a neutral accreting matter, but with other clumps or wind structures located probably farther away from the compact source. In Fig.~\\ref{fig:comp} we compare the light curves during bright flares from four SFXTs, all observed with {\\it Swift}: the two reported here from \\srcigr\\ and \\srcxte, together with the one observed from IGR~J11215--5952 \\citep{Romano2007} and from IGR~J16479-4514 \\citep{Romano2008:sfxts_paperII}. All light curves during bright flares look similar, although they were observed with a different sampling. We postpone a more quantitative comparison between the four SFXTs (duty cycle, light curve rise time and decay times) to a final paper at the end of the on-going observing campaign. In any case, it is already evident that the behaviour of this sample of SFXTs in outburst is similar, and that their bright emission extends for more than a few hours, contrary to what originally thought at the time of the discovery of this new class of sources \\citep[e.g., ][]{Sguera2005}. The wide band spectra during outbursts display high energy cut-offs (assuming the model with a power-law modified at high energy by a cutoft, {\\sc highecut} in XSPEC), although differently constrained in the two sources: in \\srcigr\\ it is at 18$\\pm{2}$~keV, in \\srcxte\\ it lies below 13~keV. These cut-off ranges are fully consistent with a neutron star magnetic field, B, of about a 2--3$\\times$10$^{12}$~G in the case of \\src\\ and of less than about 2$\\times$10$^{12}$~G for \\srcxte\\ \\citep{Coburn2002}. These estimates, although not based on a direct measurement of the magnetic field (which would be possible only in the case of detection of cyclotron lines), are already difficult to explain in the framework of the magnetar model recently proposed by \\citet{Bozzo2008}, where the magnetic field is at a level of 10$^{14}$~G. The same is true for other two SFXTs, IGR~J11215--5952 \\citep{Sidoli2007} and IGR~J16479--4514 \\citep{Romano2008:sfxts_paperII}. {\\it Facilities:} \\facility{{\\it Swift}}." }, "0808/0808.1878_arXiv.txt": { "abstract": "s{The NuMoon project uses the Westerbork Synthesis Radio Telescope to search for short radio pulses from the Moon. These pulses are created when an ultra high energy cosmic ray or neutrino initiates a particle cascade inside the Moon's regolith. The cascade has a negative charge excess and moves faster than the local speed of light, which causes coherent Cherenkov radiation to be emitted. With 100 hours of data, a limit on the neutrino flux can be set that is an order of magnitude better than the current one (based on FORTE). We present an analysis of the first 10 hours of data. } ", "introduction": "Above the Greisen-Zatsepin-Kuzmin (GZK) energy of $6\\cdot 10^{19}$~eV, cosmic rays (CRs) can interact with the cosmic microwave background photons, losing energy and producing pions when traversing distances of the order of 10 Mpc \\cite{g66,zk66}. Recent results of the Pierre Auger Observatory have confirmed a steepening in the cosmic ray spectrum at the GZK energy \\cite{Y07}. The pions will produce ultrahigh energy (UHE) neutrinos through weak decay. Observing these neutrinos is of great scientific interest as their arrival direction points back to sources at distances larger than 10 Mpc, in contrast to charged particles that are deflected in (extra-)galactic magnetic fields. In addition, while cosmic rays from faraway sources pile up at the GZK energy, information about the cosmic ray spectrum at the source is conserved in the GZK neutrino flux. Other possible sources of UHE neutrinos are decaying supermassive particles, such as magnetic monopoles or topological defects. This class of models is refered to as top-down (TD) models (see for example Stanev \\cite{s04} for a review). Because of their small interaction cross section and low flux, the detection of cosmic neutrinos calls for extremely large detectors. Assuming the Waxman-Bahcall flux \\cite{wb01}, even at GeV energies the flux is not higher than a few tens of neutrinos per km$^2$ per year. Kilometer-scale detectors are not easily built but can be found in nature. For example, interaction of neutrinos in ice or water can be detected by the Cherenkov light produced by the lepton track or cascade. The now half finished IceCube detector \\cite{icecube} will cover a km$^3$ volume of South Pole ice with PMTs, while Antares \\cite{antares} and its successor KM3NET \\cite{KM3NET} exploit the same technique in the Mediterranean sea. Even larger volumes can be covered by observing large detector masses from a distance. The ANITA balloon mission \\cite{anita} monitors an area of a million km$^2$ of South Pole ice from an altitude of $\\sim 37$~km and the FORTE satellite \\cite{forte} can pick up radio signals coming from the Greenland ice mass. Alternatively, cosmic ray experiments like the Pierre Auger Observatory can possibly distinguish cosmic ray induced air showers from neutrino induced cascades at very high declinations where the atmosphere is thickest and only neutrinos can interact close to the detector. For an even larger detector volume one can turn to the Moon. The negative charge excess of a particle shower inside a dense medium will cause the emission of coherent Cherenkov radiation in a process known as the Askaryan effect \\cite{a62}. This emission mechanism has been experimentally verified at accelerators \\cite{s01,g00} and extensive calculations have been performed to quantify the effect \\cite{zhs92,az97}. The idea to observe this type of emission from the Moon with radio telescopes was first proposed by Dagkesamanskii and Zheleznyk \\cite{dz89} and the first experimental endeavours in this direction were carried out with the Parkes telescope \\cite{parkes} and at Goldstone (GLUE) \\cite{glue}. The NuMoon project uses the Westerbork Synthesis Radio Telescope (WSRT) to watch for the same flashes but at lower frequency, which has the distinct advantage that radio pulses have a much higher chance of reaching the observer, which will be explained in the next section. ", "conclusions": "We observe at a frequency window that offers an optimal sensitivity to lunar pulses. At the same time, for these low frequencies the effects of uncertainties concerning the lunar surface and interior are small. Because of the large spread in emission angle, we expect no systematic effect from surface irregularities. The detection efficiency is also largely independent from details in the structure of the (sub-)regolith \\cite{scholten}. A possible complication is the limited dynamic range of the WSRT data. The individual dishes give a 2-bit signal, so the total dynamic range is $3N+1$ for $N$ dishes. When there is a lot of RFI background, as much as 90\\% of the total received power can be in narrow RFI lines. After background reduction the dynamic range is suppressed and detection of pulses may become impossible. Another effect of the limited dynamic range is that large pulses will be cut off to the maximum allowed value, potentially lowering the signal-to-noise. Here, the atmospheric dispersion contributes positively by spreading the power over more timebins. Currently we are simulating the detection efficiency for various radio backgrounds and atmospheric conditions. The error on the TEC value, used for dedispersion, is also taken into account." }, "0808/0808.0279_arXiv.txt": { "abstract": "{Using the atmosphere as a detector volume, Imaging Air Cherenkov Telescopes (IACTs) depend highly on the properties and the condition of the air mass above the telescope. On the Canary Island of La Palma, where the Major Atmospheric Gamma-ray Imaging Cherenkov telescope (MAGIC) is situated, the Saharan Air Layer (SAL) can cause strong atmospheric absorption affecting the data quality and resulting in a reduced gamma flux.}{To correlate IACT data with other measurements, e.g. long-term monitoring or Multi-Wavelength (MWL) studies, an accurate flux determination is mandatory. Therefore, a method to correct the data for the effect of the SAL is required.}{Three different measurements of the atmospheric absorption are performed on La Palma. From the determined transmission, a correction factor is calculated and applied to the MAGIC data.} {The different transmission measurements from optical and IACT data provide comparable results. MAGIC data of PG\\,1553+113, taken during a MWL campaign in July 2006, were analyzed using the presented method, providing a corrected flux measurement for the study of the spectral energy distribution of the source.}{} ", "introduction": "The Earth's atmosphere plays an important role in all ground-based observations because it is traversed by light from all objects observed. Scattering and absorption of light by the atmosphere is variable and reduces the flux measured at the telescope. On the Canary Island of La Palma, very strong atmospheric absorption can occur, in case the meteorologic phenomenon Saharan Air Layer (SAL) takes place. A warm air mass, consisting of mineral dust, travels from the Sahara westward; this becomes the so-called SAL, when it is undercut by a cool and moist air layer from the sea. Located between 1.5\\,km and 5.5\\,km above sea level (a.s.l.), the SAL, also known as Calima, can extend above the North Atlantic Ocean as far as the United States and the western Caribbean Sea \\citep{sal}. Variable absorption can cause apparent flux variations, and the absolute flux level is in all cases reduced. To to correct for this reduction, the atmospheric absorption is measured. On La Palma, three measurements are available. The Carlsberg Meridian Telescope (CMT) provides nightly values of extinction. From the optical data of the KVA and also from the MAGIC data itself, the atmospheric transmission can be determined. The results of these measurements are discussed and compared for the nights between July 15${\\rm ^{th}}$ 2006 and July 28${\\rm ^{th}}$ 2006. A correction method was developed for the MAGIC telescope: From the absorption measurements, a factor is calculated and used in the analysis to correct for the effect of SAL. ", "conclusions": "The different measurements of the atmospheric transmission from stellar photometry (KVA, CMT) and IACT data (MAGIC) agree well for nights with stable extinction. Since more than 90\\,\\% of the shower light originates above the SAL, the measurements can be used directly to correct the IACT data. By applying a correction factor in the calibration, the data of PG\\,1553+113 taken during the MWL campaign in July 2006 for example were corrected successfully. This shows that the loss of observation time due to atmospheric absorption can be reduced. For additional checks, affected data of a steady source are required. When such data become available within scheduled MAGIC observations, the presented method can be tested further. The correction should not be applied blindly to data affected by atmospheric absorption of another kind. For those, the vertical distribution of the atmospheric conditions should be studied e.g.\\ with a Lidar (light detection and ranging), i.e.\\ a device shooting a laser beam into the atmosphere and measuring the backscattered light. The possible use of information provided by a Lidar, however, still has to be investigated. In cases of strong absorption, i.e.\\ above 40\\,\\%, the method is limited by the fluctuations in the night-sky background light and the effect cannot be corrected completely." }, "0808/0808.2536_arXiv.txt": { "abstract": "The key contribution of the discovery of nuclear-powered pulsations from the accretion-powered millisecond pulsars (AMPs) has been the establishment of burst oscillation frequency as a reliable proxy for stellar spin rate. This has doubled the sample of rapidly-rotating accreting neutron stars and revealed the unexpected absence of any stars rotating near the break-up limit. The resulting `braking problem' is now a major concern for theorists, particularly given the possible role of gravitational wave emission in limiting spin. This, however, is not the only area where burst oscillations from the AMPs are having an impact. Burst oscillation timing is developing into a promising technique for verifying the level of spin variability in the AMPs (a topic of considerable debate). These sources also provide unique input to our efforts to understand the still-elusive burst oscillation mechanism. This is because they are the only stars where we can reliably gauge the role of uneven fuel deposition and, of course, the magnetic field. ", "introduction": "`History', as George Orwell once noted, `is written by the winners' \\citep{orw44} - or, in this case, by the workshop hosts. When we gathered in Amsterdam in April 2008, it was ostensibly to celebrate ten years since the discovery of the first Accreting Millisecond X-ray Pulsar (AMXP). We were, however, a full two years too late. For the first AMXP was not SAX J1808.4-3658 \\citep{wij98}, but rather the far less well-known 4U 1728-34 \\citep{str96b}. How on earth, you might ask, could such a slip go unnoticed? The trick, of course, lies in the terminology. Most astronomers (the author included) tend to think of the AMXPs as comprising only the {\\it accretion-powered} millisecond pulsars (AMPs), forgetting the equally large class of {\\it nuclear-powered} millisecond pulsars (NMPs) - the burst oscillation sources. Most of this volume focuses on the AMPs, where persistent pulsations are generated as accreting material is channeled by the magnetic field onto magnetic polar caps that are offset from the rotational poles. The NMPs, by contrast, show pulsations during Type I X-ray bursts (thermonuclear explosions on the stellar surface caused by rapid unstable burning of accreted material). The cause of the brightness asymmetry in the NMPs remains an open question \\citep{str06, gal08}, and to do full justice to NMP phenomenology would merit a much longer discussion. In this article, however, I will focus on the small set of NMPs that are also AMPs. These rare objects provide a unique insight into many current problems in neutron star astrophysics because, as suggested by my title, they are lighthouses with two different light sources. The accretion-powered pulsations tell us how the material arrives on the stellar surface, while the nuclear-powered pulsations tell us what happens once it gets there. Section \\ref{data} provides a brief overview of the relevant observational results. The bulk of the review focuses, however, on the astrophysical questions where these sources have made or are making a major contribution to our understanding. These include the spin distribution of AMXPs, torque modeling, and the burst oscillation mechanism. ", "conclusions": "\\label{conc} The discovery of nuclear-powered pulsations from the AMPs has cemented the link between burst oscillation frequency and spin frequency. In addition to confirming the absence of rapid rotators (now a major problem for evolutionary models), this has imposed the strongest single constraint on candidate burst oscillation mechanisms. Despite this huge clue, the mechanism remains elusive: but continuing analysis of the AMPs is providing tantalising evidence that is driving the development of new theoretical models." }, "0808/0808.2193_arXiv.txt": { "abstract": "We report on our early photometric and spectroscopic observations of the extremely luminous Type II supernova (SN) 2008es. With an observed peak optical magnitude of $m_V = 17.8$ and at a redshift $z = 0.213$, \\sn\\ had a peak absolute magnitude of $M_V$ = $-22.3$, making it the second most luminous SN ever observed. The photometric evolution of \\sn\\ exhibits a fast decline rate ($\\sim$0.042 mag d$^{-1}$), similar to the extremely luminous Type II-L SN~2005ap. We show that \\sn\\ spectroscopically resembles the luminous Type II-L SN~1979C. Although the spectra of \\sn\\ lack the narrow and intermediate-width line emission typically associated with the interaction of a SN with the circumstellar medium of its progenitor star, we argue that the extreme luminosity of \\sn\\ is powered via strong interaction with a dense, optically thick circumstellar medium. The integrated bolometric luminosity of \\sn\\ yields a total radiated energy at ultraviolet and optical wavelengths of $\\ga$10$^{51}$ ergs. Finally, we examine the apparently anomalous rate at which the Texas Supernova Search has discovered rare kinds of supernovae, including the five most luminous supernovae observed to date, and find that their results are consistent with those of other modern SN searches. ", "introduction": "Wide-field synoptic optical imaging surveys are continuing to probe the parameter space of time-variable phenomena with increasing depth and temporal coverage \\citepeg{2004ApJ...611..418B,2006MNRAS.371.1681M,2008MNRAS.386..887B}, unveiling a variety of transients ranging from the common \\citep{2008ApJ...682.1205R} to the unexplained (e.g., \\citealt{bdt+08}). Untargeted (``blind'') synoptic wide-field imaging surveys, such as the Texas Supernova Search (TSS; \\citealt{quimbyTSS}) conducted with the ROTSE-III 0.45-m telescope \\citep{akerlof03}, have uncovered a large number of rare transients, including the four most luminous supernovae (SNe) observed to date: SN~2005ap \\citep{quimby05ap}, SN~2008am \\citep{ATEL.1389}, SN~2006gy \\citep{ofek06gy, smith07-2006gy, smith08-2006gy}, and SN~2006tf \\citep{smith06tf}. Observations of these very luminous events are starting to allow the detailed physical study of the extrema in core-collapse SNe. They appear to be powered in part by their interaction with a highly dense circumstellar medium (CSM; see \\citealt{smith06tf}, and references therein), though other possibilities have been advanced. Clearly, the discovery of more such events would allow an exploration of the variety of the phenomenology as related to the diversity of progenitors and CSM. Recently, the TSS discovered yet another luminous transient on 2008 Apr. 26.23 (UT dates are used throughout this paper), which they suggested was a variable active galactic nucleus at a redshift $z = 1.02$ \\citep{ATEL.1515}. \\citet{ATEL.1524} then hypothesized that the transient was a flare from the tidal disruption of a star by a supermassive black hole. \\citet{ATEL.1576} first identified \\sn\\ as potentially an extremely luminous Type II SN (see also \\citealt{2008ATel.1578....1G}), and we later definitively confirmed this with further spectroscopic observations \\citep{ATEL.1644}; the event was assigned the name \\sn\\ by the IAU \\citep{2008CBET.1462....1C}. It is located at $\\alpha$ = 11$^h$56$^m$49.06$^s$, $\\delta$ = +54$^o$27$\\arcmin$24.77$\\arcsec$ (J2000.0). Here we present our analysis of SN 2008es, which is classified as a Type II-Linear (II-L) SN based on the observed linear (in mag) decline in the photometric light curve \\citep{barbon79,doggett85}. At $z$ = 0.213, \\sn\\ has a peak optical magnitude of $M_V = -22.3$, among SNe second only to SN~2005ap. Aside from the extreme luminosity, \\sn\\ is of great interest since detailed ultraviolet (UV) through infrared (IR) observations provide a unique opportunity to study the mass-loss properties of an evolved post-main sequence massive star via its interaction with the surrounding dense CSM. A similar analysis of \\sn\\ has been presented by \\citet{gezari08}. The outline of this paper is as follows. We present our observations in $\\S$2, and the photometric and spectroscopic analyses of this and public (NASA) data in $\\S$3 and $\\S$4, respectively. A discussion is given in $\\S$5, and our conclusions are summarized in $\\S$6. Throughout this paper we adopt a concordance cosmology of $H_0$ = 70 \\kms\\ Mpc$^{-1}$, $\\Omega_{\\rm M} = 0.3$, and $\\Omega_\\Lambda = 0.7$. ", "conclusions": "We have reported on our early observations of \\sn, which at a peak optical magnitude of $M_V = -22.3$ is the second most luminous SN ever observed. We argue that the extreme luminosity of this SN was likely powered via strong interaction with a dense CSM, and that the steep decline in the light curve, 0.042 mag d$^{-1}$, indicates that the radioactive decay of $^{56}$Co is likely not the dominant source of energy for this SN. Integration of the bolometric light curve of \\sn\\ yields a total radiated energy output of $\\ga$ 10$^{51}$ ergs. The optical spectra of \\sn\\ resemble those of the luminous SN~1979C, but with an unexplained increase in the velocity of the H$\\beta$ absorption minimum over time. We also examined the rate of discovery of extremely luminous SNe by the Texas Supernova Search and find that their discovery of the five most luminous observed SNe in the past four years is probably not a fluke; several more such detections are expected in the coming years. Finally, what behavior can we expect from \\sn\\ at late times, roughly 1 yr or more after discovery? Regardless of whether the peak luminosity was powered by radioactive decay or optically thick CSM interaction (see \\citealt{smith08-2006gy}), a SN can be powered by strong CSM interaction at late times if the progenitor had a sufficiently high mass-loss rate in the centuries before exploding. We have seen examples of both: SN~2006tf had strong H$\\alpha$ emission indicative of ongoing CSM interaction at late times \\citep{smith06tf}, whereas SN~2006gy did not \\citep{smith08-2006gy}. SN~1979C was less luminous at peak than those two SNe, but it has been studied for three decades because its late-time CSM interaction is powering ongoing emission in the radio, optical, and X-rays \\citep{weiler81,fesen93, immler98}. With such an extraordinarily high peak luminosity, a late-time IR echo such as that seen in SN~2006gy \\citep{smith08-2006gy} is also likely if \\sn\\ has dust waiting at a radius of $\\sim$0.3 pc. Alternatively, if \\sn\\ were powered in whole or in part by radioactivity, a large mass of $^{56}$Ni should be evident in the late-time decline rate." }, "0808/0808.2470_arXiv.txt": { "abstract": "We present observations of \\caii, \\znii\\, and \\crii\\ absorption lines in 16 damped Lyman alpha systems (DLA) and six subDLAs at redshifts $0.6 < z_{\\rm abs} < 1.3$, obtained for the dual purposes of: (i) clarifying the relationship between DLAs and absorption systems selected from their strong \\caii\\ lines, and (ii) increasing the still limited sample of Zn and Cr abundance determinations in this redshift range. We find only partial overlap between current samples of intermediate redshift DLAs (which are drawn from magnitude limited surveys) and strong \\caii\\ absorbers: approximately 25 per cent of known DLAs at these redshifts have an associated Ca\\,{\\sc ii}\\,$\\lambda 3935$ line with a rest-frame equivalent width greater than 0.35\\,\\AA, the threshold of the Sloan Digital Sky Survey sample assembled by Wild and her collaborators. The lack of the strongest \\caii\\ systems (with equivalent widths greater than 0.5\\,\\AA) is consistent with these authors' conclusion that such absorbers are often missed in current DLA surveys because they redden and dim the light of the background QSOs. We rule out the suggestion that strong \\caii\\ absorption is associated exclusively with the highest column density DLAs. Furthermore, we find no correlation between the strength of the \\caii\\ lines and either the metallicity or degree of depletion of refractory elements, although the strongest \\caii\\ absorber in our sample is also the most metal-rich DLA yet discovered, with [Zn/H] $\\simeq$ solar. We conclude that a complex mix of parameters must determine the strengths of the \\caii\\ lines, including the density of particles and ultraviolet photons in the interstellar media of the galaxies hosting the DLAs. We find tentative evidence (given the small size of our sample) that strong \\caii\\ systems may preferentially sample regions of high gas density, perhaps akin to the DLAs exhibiting molecular hydrogen absorption at redshifts $z > 2$. If this connection is confirmed, strong \\caii\\ absorbers would trace possibly metal-rich, H$_2$-bearing columns of cool, dense gas at distances up to tens of kpc from normal galaxies. ", "introduction": "\\label{Sec:Int} Quasar absorption lines are a powerful tool for the study of gaseous galactic structures over cosmic time, providing clues to the nature and evolution of galaxies and galactic halos. The large columns of neutral hydrogen gas traced by damped Lyman alpha (DLA) and subDLA systems (conventionally defined to have \\hi\\ column densities $N$(H\\,{\\sc i}) $\\ge 10^{20.3}$\\,\\acm\\ and $10^{19} \\le N$(H\\,{\\sc i})\\,$ < 10^{20.3}$\\,\\acm\\ respectively) are of particular interest (Rao 2005; Wolfe, Gawiser, \\& Prochaska 2005), as they account for most of the neutral gas in the Universe and trace galaxies which are often difficult to detect directly. Thus, appropriate to their importance, much effort has been expended to understand the nature and incidence of the gas traced by DLAs and subDLAs. While DLAs do contain metals (Wolfe et al.\\ 1986; Meyer, Welty, \\& York 1989; Pettini, Boksenberg, \\& Hunstead 1990) and ionised gas (Fox et al. 2007a), they are characterized by low metallicities (Kulkarni et al. 2005; Akerman et al. 2005; Prochaska et al. 2007) and large neutral fractions (Viegas 1995; Vladilo et al. 2001). Little, if any, evolution is detected in the co-moving mass density of \\hi, $\\Omega_{\\rm H\\,I}$, from $z \\simeq 4$ to $z \\approx 0.5$ (Rao, Turnshek, \\& Nestor, 2006; Lah et al.\\ 2007; although see Prochaska, Herbert-Fort, \\& Wolfe 2005) and of metals (Kulkarni et al. 2005). This is perhaps surprising, considering the putative role of DLAs as the reservoir of fuel for star formation. In fact, recent work (Wolfe \\& Chen 2006; Wild, Hewett, \\& Pettini 2007) has shown that \\textit{in situ} star formation in DLAs is significantly less than expected on the basis of the present-day Kennicutt-Schmidt law (Kennicutt 1998), considering their large \\hi\\ surface densities. One possible explanation for the apparent lack of evolution in $\\Omega_{\\rm H\\,I}$ is that a significant fraction of DLAs with high star formation rates or metallicities are missed from traditional magnitude limited optical QSO surveys due to chromatic extinction from large columns of dust. This possibility, however, is not supported by surveys of DLAs in radio-selected QSOs (Ellison et al. 2001; Jorgenson et al. 2006). SubDLAs, on the other hand, contribute a smaller fraction of the total \\hi\\ in the Universe. However, they possess larger fractions of ionised gas than DLAs, and thus the total amount of hydrogen (\\hi\\ plus \\hii) in subDLAs may be comparable to that of DLAs (Fox et al. 2007b). Furthermore, subDLAs appear to have higher metallicities than DLAs (Kulkarni et al. 2007) and thus may be a large repository of the metals produced at high redshifts (see Pettini 2006). While DLAs and subDLAs both contain large amounts of hydrogen, they are likely to arise in different regions within galaxies and to trace different environments and properties. Strong metal lines of Mg\\,{\\sc ii}\\,$\\lambda\\lambda 2796, 2804$ and \\feii\\,$\\lambda\\lambda 2587, 2600$ are ubiquitous in (sub)DLA systems. Their large oscillator strengths and near-ultraviolet (UV) rest wavelengths make them ideal for tracing large columns of neutral gas at redshifts $z \\la 1.7$, where the Ly$\\alpha$ line falls below the atmospheric cut-off at 3200\\,\\AA\\ and therefore requires observations from space. In this context, systems selected via strong Ca\\,{\\sc ii}\\,$\\lambda\\lambda 3935, 3970$ absorption have recently garnered attention (Wild \\& Hewett 2005). Extensive literature exists on the Ca\\,{\\sc ii} doublet in the interstellar media of the Milky Way and nearby galaxies, since these lines can be studied from the ground even at redshift $z=0$ (e.g. Marshall \\& Hobbs 1972; Welty, Morton, \\& Hobbs 1996 and references therein). This body of work has shown that in local interstellar environments Ca is a strongly refractory element, exhibiting some of the highest depletion factors (Savage \\& Sembach 1996). Furthermore, since the ionisation potential of Ca\\,{\\sc ii} is lower than that of hydrogen, Ca\\,{\\sc ii} is not the dominant ionisation stage of Ca in H\\,{\\sc i} regions, unlike Mg\\,{\\sc ii} and Fe\\,{\\sc ii}. Rather, the balance between Ca\\,{\\sc ii} and Ca\\,{\\sc iii} depends on the details of the physical conditions in the gas, primarily on temperature and on the densities of particles and far-UV photons. Therefore, while all DLAs exhibit strong \\mgii\\ and \\feii\\ lines, the same is not true of \\caii. Wild, Hewett, \\& Pettini (2006; WHP06) recently conducted a survey for strong \\caii\\ systems, with $\\lambda 3935$ rest equivalent width \\wca\\,$\\geq 0.35$\\,\\AA, in QSO spectra from the Sloan Digital Sky Survey (SDSS). Based on the strengths of Zn\\,{\\sc ii}\\ and Cr\\,{\\sc ii}\\ lines, WHP06 concluded that, on average, strong Ca\\,{\\sc ii} absorbers have \\nhi\\ values above the DLA limit. They also showed that these absorbers are, on average, dustier than typical DLAs, introducing a reddening of $\\langle E(B-V)\\rangle \\ga 0.1$\\,magnitudes in the spectra of the background QSOs, compared to $\\langle E(B-V)\\rangle \\la 0.02$ for the general population of SDSS DLAs (WHP06; Vladilo, Prochaska, \\& Wolfe 2008). This level of obscuration of the background QSOs is significant, as it causes flux-limited and/or colour-selected surveys to underestimate the numbers of strong Ca\\,{\\sc ii} systems. If \\caii-strong DLAs are being missed, and if they are preferentially metal-rich relative to DLAs with weaker \\caii\\ absorption, then this class of absorption system may contain a previously overlooked repository of both neutral gas and metals. In any case, if \\caii\\ systems select a population of absorbers with unique properties, they could hold the key to a better understanding of the detailed physical properties of DLAs and subDLAs. Since the survey by WHP06, efforts have been devoted to better understanding the link between strong Ca\\,{\\sc ii} systems and galaxies. Using $K$-band imaging, Hewett \\& Wild (2007) reported an excess of luminous galaxies close to the sightlines to 30 QSOs with intervening Ca\\,{\\sc ii} systems at redshifts $0.7 < z < 1.2$. The galaxies appear to exhibit a luminosity-dependent cross section for \\caii\\ absorption and have a mean impact parameter of $\\approx 25$\\,kpc. Together with the incidence of absorbers from WHP06, this result implies a halo volume filling factor of $\\sim 10$ per cent out to $\\approx 35$\\,kpc from the absorbing galaxies. At lower redshifts, Zych et al. (2007) were able to associate luminous, metal-rich, star-forming spiral galaxies with four out of five strong Ca\\,{\\sc ii} absorbers at $z < 0.5$. Although these studies represent progress in uncovering the nature of \\caii\\ absorbers, ignorance of \\nhi\\ in such systems limits the direct insight they provide into DLAs and subDLAs in general. While it seems very likely that \\caii\\ absorbers select large columns of neutral gas, direct measurements of \\nhi\\ are currently available for very few \\caii\\ systems. The reason is simple: by the time the SDSS data highlighted the potential importance of absorption systems selected via Ca\\,{\\sc ii}, no space-borne UV spectrograph was in operation to measure the associated column densities of H\\,{\\sc i}. This unfortunate state of affairs is about to change with the forthcoming installation of the Cosmic Origins Spectrograph (COS) on the refurbished \\textit{Hubble Space Telescope} (\\textit{HST}). In the meantime, the most immediate approach to this problem is to perform the complementary experiment of measuring the strength of \\caii\\ absorption in known DLAs at redshifts $z \\simlt 1.3$, where the \\caii\\ doublet lines are still accessible with optical spectrographs. The samples of confirmed DLAs at these intermediate redshifts have increased considerably in recent years thanks to dedicated \\textit{HST} surveys (see Rao, Turnshek \\& Nestor 2006; RTN06). In this paper, we present new observations of QSOs from the RTN06 compilation designed to: \\textit{(i)} measure the equivalent widths of the Ca\\,{\\sc ii}\\,$\\lambda \\lambda 3935, 3970$ doublet lines in order to assess the overlap of known DLAs (and a few subDLAs) with the strong Ca\\,{\\sc ii} absorber sample of WHP06; and \\textit{(ii)} simultaneously determine the metallicities and dust content of intermediate redshift DLAs via observations of the associated Zn\\,{\\sc ii} and Cr\\,{\\sc ii} absorption lines (Pettini et al. 1990), thereby adding to the still somewhat limited statistics of these measures at $z < 1.5$ (Akerman et al. 2005; Kulkarni et al. 2007). In Section~\\ref{Sec:Obs} we describe our observations and equivalent width measurements, while Section~\\ref{sec:abundances} deals with the derivation of column densities and element abundances. The main results on the \\caii\\ properties of known DLAs at intermediate redshifts are presented in Section~\\ref{sec:caii_in_dlas}. We summarise our principal findings and conclusions in Section~\\ref{Sec:Sum}. ", "conclusions": "\\label{Sec:Sum} We have used the WHT telescope on La Palma to record at intermediate resolution the optical spectra of QSOs known to lie behind DLAs and subDLAs in the redshift range $0.6 < z_{\\rm abs} < 1.3$. Our principal aims were twofold: (i) to measure the strength of the Ca\\,{\\sc ii}\\,$\\lambda \\lambda 3935, 3970$ absorption lines in known DLAs, with a view to establishing to what extent DLAs and strong Ca\\,{\\sc ii} absorbers overlap; and (ii) to measure the \\znii\\,$\\lambda\\lambda 2026, 2063$ and \\crii\\,$\\lambda\\lambda 2056, 2062, 2066$ multiplets in the (sub)DLAs, thereby increasing the number of metallicity and depletion determinations in (sub)DLAs at low-redshift. By design, these complementary measurements allow us to test whether the strength of Ca\\,{\\sc ii} absorption is related to the metallicity of the gas, as measured by the abundance of Zn, and/or to the degree of dust depletion of refractory elements (both Ca and Cr are readily incorporated onto dust, whereas Zn is not). From consideration of such measurements in a sample of 16 DLAs and six subDLAs, augmented by similar data for three DLAs and four subDLAs from published studies, we reach the following conclusions. \\begin{itemize} \\item[1.] The new measurements presented here strengthen previous conclusions on the metallicities and dust depletions in DLAs at $z < 1.3$. We point out that, after accounting for the fact that at these redshifts DLAs have mostly been selected from Mg\\,{\\sc ii}-strong systems -- and may thus be biased in favour of metal-rich absorbers -- the already weak redshift evolution of the mean DLA metallicity is reduced further: [$\\langle {\\rm Zn/H}\\rangle$] is only a factor of about two greater at $0.6 < z < 1.3$ than at higher redshifts. We also confirm the dependence of dust depletion on metallicity uncovered by earlier studies. \\item[2.] Most DLAs exhibit detectable Ca\\,{\\sc ii}\\,$\\lambda 3935$ absorption in our spectra. Our new data indicate that selecting absorption systems above the threshold \\wca\\,$ = 0.15$\\,\\AA\\ isolates systems with \\nhi\\,$\\ga 2 \\times 10^{20}$\\,cm$^{-2}$ (although with some contamination from subDLAs, as evidenced by detections reported in the literature). However, while it seems likely that strong Ca\\,{\\sc ii} absorption is associated with DLAs, the converse is not true: approximately one third of the DLAs in our sample have \\wca\\,$< 0.15$\\,\\AA, while adopting the criterion \\wca\\,$> 0.30$\\,\\AA\\ would miss $\\simeq 3/4$ of the DLAs. \\item[3.] Only four out of the 19 DLAs in the combined sample meet the selection criterion \\wca\\,$\\geq 0.35$\\,\\AA\\ of the SDSS survey by WHP6, and only one out of 19 would be classed as a `strong' Ca\\,{\\sc ii} absorber by their definition (\\wca\\,$\\geq 0.5$\\,\\AA). Evidently, there is little overlap as yet between confirmed DLAs and the SDSS candidates identified via Ca\\,{\\sc ii}. This is not surprising, however, since -- as demonstrated by WHP06 -- dust associated with many of the strong Ca\\,{\\sc ii} absorbers dims the light from the background QSOs sufficiently for them to be excluded from the magnitude-limited surveys conducted with \\textit{HST} up to now. \\item[4.] Overall, we find no convincing correlation of the strength of Ca\\,{\\sc ii} absorption with the column density of neutral gas, the equivalent width of Mg\\,{\\sc ii}\\,$\\lambda 2796$, the metallicity of the gas (as measured from the abundance of Zn), or the degree of depletion of refractory elements (indicated by the [Cr/Zn] ratio). Despite the fact that, in the interstellar medium of the Milky Way, both Ca and Cr exhibit large and variable depletion factors from the gas-phase, DLAs with near-solar values of [Cr/Zn] do not show significantly stronger Ca\\,{\\sc ii} lines than those where most of the Cr is presumably in solid form. On the other hand, the finding that the only `strong' \\caii\\ absorber in our sample is the most metal-rich DLA yet discovered suggests that there may be some connection between the metallicity and \\caii\\ equivalent width. \\item[5.] Presumably, a more complex mix of factors than simply H\\,{\\sc i} column density, metallicity, and dust depletion, determines the value of \\wca\\ in DLAs. We have presented preliminary evidence that the `joker in the pack' may be the variable fraction of Ca which is singly ionised in H\\,{\\sc i} regions, where Ca\\,{\\sc iii} is often the dominant ion stage, depending on the densities of particles, far-UV photons, and on the temperature of the gas. With the data at hand, we exploit the ionisation balance of Ca to place constraints on the ratio of UV photons and particles. If we further assume that the interstellar radiation field in the DLAs is of the same order as that at our location within the Milky Way -- as indeed found for DLAs at higher redshifts -- we can arrive at estimates of, or limits on, the gas density $n$(H$_{\\rm TOT})$. Our modelling, using the photoionisation code {\\sc cloudy}, indicates values $n$(H$_{\\rm TOT}) \\simgt 1$\\,cm$^{-3}$ for most DLAs with \\wca\\,$ \\simgt 0.35$\\,\\AA. Furthermore, for an `average' \\caii-absorber DLA (as described by WHP06) as well as the lone `strong' (as defined by those authors) \\caii\\ system in our sample, we deduce $n$(H$_{\\rm TOT}) \\simgt 10$\\,cm$^{-3}$. Such high values are similar to those which seem to be typical of H$_2$-bearing DLAs at $z > 2$, raising the possibility that in the strong Ca\\,{\\sc ii} absorbers selected by WHP06 at $z < 1.3$ we may be seeing the lower redshift counterparts of the subset of DLAs with a measurable molecular fraction. Indeed, our models predict molecular fractions similar to those reported for relatively metal-rich DLAs at high-$z$, with the strongest \\caii\\ absorbers exhibiting values consistent with the highest reported measurements of $f_{\\rm H_2}$ in DLAs. \\end{itemize} If borne out by additional studies, the possible connection of strong \\caii\\ absorbers to metal- and relatively molecule-rich gas has intriguing consequences when considered in light of the imaging results of Hewett \\& Wild (2007). It would mean that strong \\caii\\ absorbers select gas with these properties not only in galactic disks, but also at distances of tens of kpc from relatively bright galaxies. Such a conclusion would be highly relevant to models of galactic outflows and star-formation feedback. The currently observed population of DLAs are {\\it not}, in general, strong \\caii\\ absorbers. However, if most or all strong \\caii\\ absorbers are DLAs and the true incidence of strong \\caii\\ absorbers is as high as calculated by WHP06, these intriguing systems may represent a (potentially metal-rich) subset of DLA absorbers that are being missed from the current surveys. It will be possible to make progress towards clarifying this and other questions raised by the present work with COS measurements of metal lines and \\nhi\\ in a sample of {\\it known} strong \\caii\\ absorbers. Like the rest of the UV-spectroscopy community, we look forward with anticipation to the forthcoming \\textit{HST} servicing mission that will install this much needed instrument." }, "0808/0808.0196_arXiv.txt": { "abstract": "A weakly interacting massive particle (WIMP) weighing only a few GeV has been invoked as an explanation for the signal from the DAMA/LIBRA experiment. We show that the data from DAMA/LIBRA are now powerful enough to strongly constrain the properties of any putative WIMP. Accounting for the detailed recoil spectrum, a light WIMP with a Maxwellian velocity distribution and a spin-independent (SI) interaction cannot account for the data. Even neglecting the spectrum, much of the parameter space is excluded by limits from the DAMA unmodulated signal at low energies. Significant modifications to the astrophysics or particle physics can open light mass windows. ", "introduction": "Introduction} The DAMA/LIBRA NaI(Tl) scintillation experiment \\cite{Bernabei:2008yi} has used the annual modulation technique \\cite{Drukier:1986tm,Freese:1987wu} to search for dark matter (DM). They now find a modulation of over 8$\\sigma$ with a period and phase consistent with a DM signal. However, in some models of DM, it is not trivial to square this positive result with the null results from other direct detection experiments. Recent investigations \\cite{Foot:2008nw,Feng:2008dz,Petriello:2008jj,Bottino:2008mf}, updating the discussion of \\cite{Gondolo:2005hh} note that light DM, with mass of a few GeV might reconcile the DAMA/LIBRA data with constraints from other experiments, e.g., \\cite{CDMS,XENON}. In addition, a DM candidate with mass $\\sim$ GeV is tantalizing -- it might give insight into the ratio of the DM density to the density of baryons. In this note, we point out that the statistics of the DAMA/LIBRA data are now sufficiently powerful that an explanation of the DAMA/LIBRA data must now go beyond just fitting the overall modulation rate. Additional self-consistency checks on the light WIMP scenario are now possible. We discuss two such checks. First, DAMA/LIBRA has now measured the modulation rate as a function of the observed recoil energy. This spectrum contains valuable information. Simple kinematics indicate that the spectra of nuclei recoiling against a WIMP are sensitive to the mass of the WIMP. Thus, fitting the observed energy spectrum constrains the mass of the candidate WIMP particle. Another constraint can be derived by looking at the total (unmodulated) rate of observed events with low energy recoils. Some WIMP candidates will provide more events than the {\\it total} number of observed events at low energies, despite a presumably sizable background. These two constraints are powerful probes of the light WIMP region, and effectively exclude this interpretation if a Maxwellian velocity distribution with standard parameters is assumed. Modifications to this assumption, in particular DM streams, can open up small regions of allowed parameter space. ", "conclusions": "Conclusions} The recent update from the DAMA/LIBRA collaboration has added significance to the modulated signature. Detailed spectral data should now be considered for any model proposed to explain the signal. As it appears that modulation extends up to at least 5 keVee, the spectrum places very strong constraints on the properties of a standard WIMP in a MB velocity distribution. Additionally, because the signal from a standard WIMP rises very rapidly at low recoil energies, one must insure that the {\\em unmodulated} signal predicted does not exceed the observed levels at DAMA/LIBRA. Fitting to the spectral data, one finds two regions for a standard WIMP in a Maxwellian halo which explain the spectrum. The high mass region conflicts with other direct detection experiments and the unmodulated low-energy rate at DAMA/LIBRA. The light mass region gives a poorer fit to the spectrum, but is still acceptable and is largely consistent with limits from the DAMA unmodulated signal. However, 90\\% CL limits from CDMS-Si and XENON-10 exclude this region. Modifications to astrophysics or particle physics can open light mass windows. Streams can open regions in the $m_\\chi \\sim 2 -5\\; \\gev$ range, although for very particular choices of parameters. Inelasticity can broaden areas at higher masses, but faces model-building challenges. Future direct detection experiments should push to explore both of these regions." }, "0808/0808.3476_arXiv.txt": { "abstract": "We consider the role of deuterium as a potential marker of location and ambient conditions during the formation of small bodies in our Solar system. We concentrate in particular on the formation of the regular icy satellites of Jupiter and the other giant planets, but include a discussion of the implications for the Trojan asteroids and the irregular satellites. We examine in detail the formation of regular planetary satellites within the paradigm of a circum-Jovian subnebula. Particular attention is paid to the two extreme potential subnebulae -- ``hot'' and ``cold''. In particular, we show that, for the case of the \"hot\" subnebula model, the D:H ratio in water ice measured from the regular satellites would be expected to be near-Solar. In contrast, satellites which formed in a ``cold'' subnebula would be expected to display a D:H ratio that is distinctly over-Solar. We then compare the results obtained with the enrichment regimes which could be expected for other families of icy small bodies in the outer Solar system -- the Trojan asteroids and the irregular satellites. In doing so, we demonstrate how measurements by Laplace, the James Webb Space Telescope, HERSCHEL and ALMA will play an important role in determining the true formation locations and mechanisms of these objects. ", "introduction": "\\label{intro} The physical and chemical characteristics of the proto-planetary nebula from which our solar system formed can be inferred through the analysis of ``primitive'' objects such as meteorites, comets, and the giant planets themselves. We can obtain useful constraints on these processes by examining the degree to which fossil deuterium contained within the water in some of these objects is enriched when compared to the protosolar abundance, which can be measured in various objects within the Solar system. Within the Solar nebula, the main reservoir of deuterium was molecular hydrogen. However, isotopic exchange occurred between this hydrogen and other deuterated species, resulting in the formation of secondary reservoirs of deuterium. As a result of the high cosmic abundance of oxygen, the most important secondary reservoir in the nebula is water (HDO), either in gaseous or solid phase. Calculations of the temporal and radial evolution of the D:H ratio in the primitive nebula (Drouart et al., 1999; Mousis et al., 2000) have been performed in order to reproduce existing data on comets (Balsiger et al., 1995, Eberhardt et al., 1995; Bockel{\\'e}e-Morvan et al., 1998; Meier et al., 1998; see Horner et al., 2007), and measurements taken from meteorites (Deloule et al., 1998). One particularly interesting result of these calculations is that the D:H ratio in water ice produced in the nebula varies by a significant amount as a function of the distance from the Sun at which the ice was formed. This variation can be seen clearly in the compilation of measurements given by Drouart et al. (1999), in Fig.1 of that work. Such results led Horner et al. (2007) to discuss how measurements of the D:H ratio in cometary bodies might prove helpful in answering the question of whereabouts in the Solar system the different cometary populations had formed. The study of D:H extends beyond the study of cometary bodies, however. In order to consider the measured D:H enrichment within an object to be the direct result of its formation, it is clear that the deuterated reservoir incorporated within that object must have undergone little or no alteration since its formation within the nebula. Therefore, despite the fact that deuterium enhancement has been measured on the Earth, Venus and Mars, it is obvious that we cannot consider the resulting value to be primordial. Indeed, the value measured on the Earth could be the result of the combination of volatile material from a number of sources (Dauphas et al., 2000). Futhermore, the values measured on Venus and Mars are believed to be the result of strong atmospheric fractionation, which has occurred throughout the history of the planets (Donahue, 1999; Bertaux \\& Montmessin, 2001). Even further from the Sun, observations of Titan reveal an unexpectedly high D:H ratio within the methane of the satellite's atmosphere (B{\\'e}zard et al., 2007). The cause of this enhancement is still under some debate. It could be the result of isotopic exchange between methane and molecular hydrogen within the early Solar nebula, prior the formation of the icy planetesimals that were ultimately accreted by Titan (e.g. Mousis et al., 2002a). Alternatively, it could result from a photochemical process occurring within the atmosphere of the satellite (Lunine et al., 1999). A number of ambitious projects, such as the proposed Laplace\\footnote{http://jupiter-europa.cesr.fr/} mission and the James Webb Space Telescope (scheduled for launch in 2013), will allow us to revisit the Jovian satellite system, and provide new measurements of the satellites of the giant planet in the coming years. As such, the time seems right to revisit the question of how these satellites formed, and what effect their formation would have on the quantities that could be observed by such a mission. In light of these proposed missions, such work takes on an interesting new aspect. How would the D:H ratio in the Jovian (or Saturnian) satellites be affected by their formation? Could measurements of deuterium in these satellites help us to understand the final stages of the formation of their parent bodies? Furthermore, it is important to examine such ideas during the period over which the missions are designed, to enable the construction of instruments which are fully capable of answering the questions posed. In this work, we aim to detail the various properties which would have affected the regular satellites during their formation, highlighting how the fractionation of deuterium within their ices may represent a vital window into their formation. We also examine the benefits that observations of the degree of deuteration in the irregular satellites and the Trojan asteroids would have for our understanding of the origin of the volatiles they contain, together with presenting a discussion of how and when such observations may be made. The simple picture portrayed for the comets, although applicable for other objects which formed freely in the Solar nebula (such as the asteroids), becomes more complicated when one wishes to understand the formation of the satellites of the giant planets. For the regular satellites, it is possible that material from the Solar nebula underwent a significant amount of additional processing within the planetary sub-nebula. In section 2, we will review the main results of prior discussions and calculations involving the D:H ratio in the Solar nebula, while in section 3 we show how further processing within planetary sub-nebulae would eventually lead to a reduction in the D:H ratio incorporated in the satellites forming therein when compared to gas at an equivalent distance in the Solar nebula. In section 4, we examine the cases of the irregular planetary satellites and the Jovian Trojan asteroids, two groups of object for which the study of deuteration may help untangle the minutae of their formation. The measurement of the D:H ratio in satellites and other objects by future space missions is discussed in section 5, and, in section 6, we conclude with a summary and discussion of our ideas, along with their implications for future missions to, and observations of, the giant planets and their satellite systems. ", "conclusions": "In previous work (Horner et al., 2007), we examined the role that variations in the D:H ratio through the Solar nebula would have on cometary objects observed to originate from different reservoirs in the outer Solar system. The goal of that work was to highlight that observations of the D:H ratio in such objects could prove a cornerstone in helping our understanding of the different regions in which the various cometary populations formed. Here, we have extended our arguments to include the other populations of hydrated objects in the outer Solar system - the icy planetary satellites (both regular and irregular), and the Jovian and Neptunian Trojans. We have shown that, since these objects cover a wide range of formation scenarios, the study of the D:H incorporated in their water ice provides a useful tool to answer questions on their origin. Did the regular satellites form in a hot or cold subnebula? Do the irregular planetary satellites truly represent a captured and then shattered population of objects? If so - from where were they captured? What was the origin of the Jovian and Neptunian Trojans? We focus, in particular, on the formation of regular satellites in a circum-planetary disk of gas and dust around Jupiter -- the Jovian subnebula. Current theories which describe the formation of such satellites cover a wide range of initial conditions. However, the two extreme cases are the ``hot'' and ``cold'' subnebula models. In the former, the entire subdisk is sufficiently warm that icy material falling into the subnebula from the Solar nebula is entirely vaporised, allowing the exchange of deuterium between molecular hydrogen and water to continue in the gas phase, when the water involved would otherwise already be trapped as ice. The other extreme, the ``cold'' model, assumes that the subnebula was sufficiently cold that none of the infalling volatiles were vaporised prior to their accretion in the planetesimals which went on to form the satellites. Although the true formation scenario for the regular satellites undoubtably lies somewhere between these extremes, they allow us to constrain the behaviour of the ices which went on to form the satellites we currently observe. In the case of the ``hot'' subnebula model, we have shown that the effect of the re-vaporisation of the infalling water ice allows gas phase reactions between HDO and $H_2$ to occur, leading to the gradual depletion of deuterium within the initially D-rich water. As a result, satellites which formed in this way would be expected to exhibit D:H ratios in their ice which are close to the Solar value. By contrast, in the case where the satellites form in a ``cold'' subnebula, since vaporisation of the volatiles does not occur, the D:H ratio within the satellite ices would be significantly higher, representative of material within the Solar nebula at the heliocentric distance at which the satellite's parent planet formed. The measurement of the D:H ratios within the regular satellites therefore provides a key constraint for the discussion of their formation. Although we do not go into specific detail, the models described above are equally applicable for satellites forming in the Saturnian subnebula. In the future, new observatories and space missions will allow us to measure $f$ with a greater accuracy, and in many more objects, than ever before. HERSCHEL, ALMA, Rosetta, and the JWST will allow the measurement of deuterium in objects as diverse as KBOs, comets, and the satelites of the giant planets, while the proposed Laplace Jupiter orbiter offers a unique opportunity to measure deuterium {\\it in-situ}, through use of a dust collector/ablation system, allowing the direct comparison of the regular and irregular Jovian satellites, and offering new insights into their formation. Early discussion of the type of measurements required is vital for those involved in the planning stages of these projects, in order that appropriate instrumention is constructed to make the required measurements. Another factor which may have influenced the final value of $f$ obtained by objects (particularly the giant planets and their regular satellites) is that the giant planets are thought to have migrated over a significant distance during their formation and subsequent evolution. In the most extreme examples, it has even been suggested that Uranus and Neptune formed between the orbits of Jupiter and Saturn (Levison et al., 2004)! In these cases, it is likely that the value of $f$ shown by the regular satellites of these planets will provide a doubly useful tool in determining the true history of the outer Solar system, even though it seems likely that these satellites have been destroyed and reassembled since their formation (e.g. Banfield \\& Murray, 1992). If it is the case that Uranus and Neptune formed far closer to the Sun than their current location, then it seems likely that the material incorporated in their regular satellites could reflect their formation and migration. It must be pointed out, here, that current measurements of the $f$-value in the giant planets themselves provides no useful constraint on their formation location, due to the fact that the measured value is heavily affected by the gaseous hydrogen incorporated during their formation, in addition to water obtained from planetesimals. Given a sufficiently good interior model (e.g. Feuchtgruber et al. 1999), it may be possible to extract the native $f$-value of the accreted icy fraction of the planet, which would clearly prove very useful in the study of the planetary migration. However, the satellites currently offer a simpler solution to the problem, since they formed solely from the accretion of planetesimals. Studies of these objects could provide detailed information on their $f$-values, which can constrain their migration and formation histories. Were the satellites present prior to the migration? Did they form afterward, or during the process? Similarly, should it turn out that the ``cold subnebula'' model is the fairest representation of the formation of the regular Jovian and Saturnian satellites, then it is possible that these objects contain, trapped within their ices, a record of the migration of their parent planets through the Solar nebula. Clearly, such measurements could even be used to place constraints on the distance over which the planets migrated." }, "0808/0808.3195_arXiv.txt": { "abstract": "Electrostatic behavior of a collisionless plasma in the foot region of high Mach number perpendicular shocks is investigated through the two-dimensional linear analysis and electrostatic particle-in-cell (PIC) simulation. The simulations are double periodic and taken as a proxy for the situation in the foot. The linear analysis for relatively cold unmagnetized plasmas with a reflected proton beam shows that obliquely propagating Buneman instability is strongly excited. We also found that when the electron temperature is much higher than the proton temperature, the most unstable mode is the highly obliquely propagating ion two-stream instability excited through the resonance between ion plasma oscillations of the background protons and of the beam protons, rather than the ion acoustic instability that is dominant for parallel propagation. To investigate nonlinear behavior of the ion two-stream instability, we have made PIC simulations for the shock foot region in which the initial state satisfies the Buneman instability condition. In the first phase, electrostatic waves grow two-dimensionally by the Buneman instability to heat electrons. In the second phase, highly oblique ion two-stream instability grows to heat mainly ions. This result is in contrast to previous studies based on one-dimensional simulations, for which ion acoustic instability further heats electrons. The present result implies that overheating problem of electrons for shocks in supernova remnants is resolved by considering ion two-stream instability propagating highly obliquely to the shock normal and that multi-dimensional analysis is crucial to understand the particle heating and acceleration processes in shocks. ", "introduction": "The discovery of thermal and synchrotron X-rays from young supernova remnants (SNRs) provides the evidence that electrons are heated up to a few keV and that a portion of them are accelerated to highly relativistic energy in SNR shocks \\citep{koy95}. Because SNR shocks are collisionless, not only particle acceleration mechanisms but also electron heating mechanisms in SNR shocks are not so simple. Previous studies have given an important key to the formation mechanism of perpendicular collisionless shocks. When the Alfv${\\rm \\acute{e}}$n Mach number $M_{\\rm A}$ is larger than the critical Mach number, about 3, a perpendicular shock reflects some of the incident ions to the upstream, where a foot region forms on a spatial scale of the ion gyroradius \\citep{ler83}. The plasma in the foot region consists of incident ions and electrons and reflected ions and returning ions which are made from reflected ions and move to the shock after a gyration. As for the electron heating machanism, \\citet{pap88} proposed that when the Mach number is larger than $0.5(m_{\\rm p}/m_{\\rm e})^{1/2}\\sim 20$, incident electrons and reflected ions excite electrostatic waves by the Buneman instability \\citep{bun58} because the relative velocity between them is large compared with the electron thermal velocity. They also suggest that after electrons are heated by electrostatic waves induced by the Buneman instability, ion acoustic instability is triggered because the electron temperature becomes much higher than the proton temperature. As a result, electrons are strongly heated by the Buneman instability and the ion acoustic instability. \\citet{car88} performed a one-dimensional hybrid simulation and demonstrated that strong electron heating actually occurs. They concluded that an $M_{\\rm A}=500$ shock heats electrons by a factor of $10^5$ across the shock. This means that if the upstream electron temperature is 1 eV, the downstream electron temperature becomes 100 keV. This value of the downstream temperature is much larger than the recent observational one for SNRs, a few keV \\citep{sta06}. This discrepancy has been an open issue to be resolved for a long time. On the other hand, \\citet{shi00} and \\cite{hos02} performed one-dimensional full particle-in-cell (PIC) simulations to investigate the electron acceleration at perpendicular shocks. Their simulation solves a whole region of a collisionless perpendicular shock and makes reflected ions self-consistently by employing a small proton to electron mass ratio. Their results showed that electrons are not only heated at the foot region but also significantly accelerated by surfing acceleration mechanism. However, this acceleration is valid only for the one-dimensional case because the surfing acceleration strongly depends on the structure of the electrostatic potential. In our first paper \\citep{ohi07}, we performed two-dimensional electrostatic PIC simulations to solve for the two-dimensinal structure of the electrostatic potential excited by the Buneman instability. We employed the real mass ratio but the simulation region is limited to the foot region. Our results showed that oblique modes grow as strongly as the modes parallel to the beam direction, that the potential structure becomes two-dimensional and that no efficient surfing acceleration occurs, while electron heating occurs. Thus, the problem of electron acceleration has been back to the start again. In that paper, we concentrated on the stage of the Buneman instability and did not follow the long time scale evolution after the Buneman nstability has saturated. In this paper, we study the time evolution of electrostatic collisionless plasma instabilities in the foot region by making linear analysis and by performing two-dimensional electrostatic PIC simulation. We perform simulations with a higher resolution, a larger simulation box and a longer simulation time than in \\citet{ohi07}. Especially, we focus on the evolution after electrostatic waves excited by the Buneman instability have decayed. Our simulation substantially improves most previous works that are one-dimensional, employ an artifically small proton electron mass ratio or impose rather strong magnetic field. It is obvious that a multi-dimensional analysis is necessary as discussed above \\citep{blu60, lam74, ohi07}. Our motivation for employing the real mass ratio is as follows. For a small mass ratio, the foot region in the simulation is shorter than the realistic one and the time scale on which electrons stay in the foot region in the simulation is also shorter than the realistic one because the size of the foot region is about the ion gyroradius $m_{\\rm p}v_{\\rm d}c/eB$, where $v_{\\rm d}$ and $B$ are the drift velocity of reflected protons and the magnetic field, respectively. Because reflected ions have a large free energy, we expect that more energy is transported to electrons through collective instabilities with the realistic mass ratio in the foot region. The drift velocity is not large enough to excite electromagnetic waves, so that electrostatic waves are more important. In \\S 2 we perform linear analysis for two-dimensional electrostatic modes. In \\S 3 we describe the initial setting of the PIC simulations and numerical results, followed by a discussion in \\S 4. ", "conclusions": "Now we discuss the final outcome of the ion two-stream instability. It becomes unstable when two conditions are satisfied; one is the resonance condition, $k_{x} v_{\\rm d} \\sim \\omega_{\\rm pi}$ and the other is that the wave length be between the ion Debye length and the electron Debye length. Thus, the ion two-stream instability becomes stabilized when $T_{\\rm p} \\sim T_{\\rm e}$. At the end of our simulation, the $y$-component of electric field has not decayed still completely. So, we expect $T_{\\rm e}/T_{\\rm p} < 10$ in the final stage, and probably ions will be heated up to $T_{\\rm i} \\sim m_{\\rm e} v_{\\rm d} ^2 \\sim 4m_{\\rm e} v_{\\rm sh}^2$ at the foot region in high Mach number perpendicular shocks. Of course, because the proton temperature of the drift direction is still cold in the present simulation, we must check the isotropilazation process of ion velocity distribution by doing longtime full PIC-simulations. As for the electrons, the electron temperature in the foot region is also about $m_{\\rm e} v_{\\rm sh}^2$. In the later stage, the growth of the two-stream instability dominates over the ion acoustic instability and little electron heating occurs. Hence, as mentioned in \\citet{ohi07}, if other electron heating mechanisms do not exist, after passing the shock front, electrons will undergo the adiabatic heating and finally, the electron temperature in the downstream becomes \\begin{equation} T_{\\rm e}\\sim 4\\times \\frac{1}{2}m_{\\rm e}(2v_{\\rm sh})^2 = 0.41{\\rm keV} \\left(\\frac{v_{\\rm sh}}{0.01c}\\right)^2, \\end{equation} where we assume that the compression ratio is 4. This has a very important implication for the electron heating process of the SNR shocks. The proton temperature in the downstream is $T_{\\rm p}=3m_{\\rm p}v_{\\rm sh}^2/16$, hence the ratio of two temperatures is \\begin{equation} T_{\\rm e}/T_{\\rm p} \\sim \\frac{128}{3}\\frac{m_{\\rm e}}{m_{\\rm p}} \\sim 0.023. \\end{equation} This value is close to the observed value as long as the shock velocity $v_{\\rm sh}$ is larger than $1500{\\rm km/s}$ \\citep{ade08}. Namely, we expect that the overheating problem of electrons raised by \\citet{car88} can be solved by the ion two-stream instability. In this paper, we prepare three ion beams as the initial condition. We have also performed simulations for other initial conditions such that there exist upstream electrons, upstream and reflected protons. Electrons and reflected protons have drift velocities to satisfy the vanishing current condition. This situation corresponds to the initial phase of the shock reformation phenomena. The results turn out to be basically the same as those in the results presented in this paper. Our simulation does not include electromagnetic modes. In the linear stage of the Buneman and ion two-tream instabilities, the effects are negligible because the growth rates of electromagnetic modes are smaller than those of electrostatic modes for $v_{\\rm d} \\sim 0.01c$. In the nonlinear stage, electromagnetic modes may be important. One of the reasons is that the electrostatic waves with a wave vector almost perpendicular to the drift direction may make currents because the charge fluctuation can not be shielded by electrons in this situation, where the fluctuation scale is smaller than the electron Debye length. Consequently the current may make the magnetic field. It is an interesting speculation that the magnetic field might be amplified more rapidly by the ion two-stream instability than the ion Weibel instability. The other reason is that anisotropic ion heating caused by the highly oblique ion two-stream instability excites the Weibel instability due to ion temperature anisotropy. At the end of our simulations, the ion temperature of $x$-direction is almost the same as initial one and the ratio of the ion temperature of $y$-direction to that of $x$-diretion $T_{\\rm iy}/T_{\\rm ix}$ is about 100. These two features may lead to magnetic field amplification and accompanying particle acceleration and heating in the shock foot region. We will make full-PIC simulations in future work to investigate these issues." }, "0808/0808.1968_arXiv.txt": { "abstract": "We present the first results on galaxy metallicity evolution at z$>$3 from two projects, LSD (Lyman-break galaxies Stellar populations and Dynamics) and AMAZE (Assessing the Mass Abundance redshift Evolution). These projects use deep near-infrared spectroscopic observations of a sample of $\\sim$40 LBGs to estimate the gas-phase metallicity from the emission lines. We derive the mass-metallicity relation at z$>$3 and compare it with the same relation at lower redshift. Strong evolution from z=0 and z=2 to z=3 is observed, and this finding puts strong constrains on the models of galaxy evolution. These preliminary results show that the effective oxygen yields does not increase with stellar mass, implying that the simple outflow model does not apply at z$>$3. ", "introduction": "Metallicity is one the most important property of galaxies, and its study is able to shed light on the detail properties of galaxy formation. It is an integrated property, not related to the present level of star formation the galaxy, but rather to the whole past history of the galaxy. In particular, metallicity is sensitive to the fraction of baryonic mass already converted into stars, i.e., to the evolutive stage of the galaxy. Also, metallicity is effected by the presence of inflows and outflows, i.e., by feedback processes and the interplay between the forming galaxy and the intergalactic medium. It is well known that local galaxies follow a well-defined mass-metallicity relation, where galaxies with larger stellar mass have higher metallicities (\\cite{tremonti04,lee06}). The origin of the relation is uncertain because several effects can be, and probably are, active. It is well known that in the local universe starburst galaxies eject a significant fraction of metal-enriched gas into the intergalactic medium because of the energetic feedback from exploding SNe, both core-collapse and, possibly, type Ia (\\cite{mannucci06}). Outflows are expected to be more important in low-mass galaxies, where the gravitational potential is lower and a smaller fraction of gas is retained. As a consequence, higher mass galaxies are expected to be more metal rich (see, for example, \\cite{edmunds90,garnett02}). A second possibility is related to the well know effect of ``downsizing'' (e.g., \\cite{cowie96}), i.e., lower-mass galaxies form their stars later and on longer time scales. At a given time, lower mass galaxies have formed a smaller fraction of their stars, therefore are expected to show lower metallicities. Other possibilities exist, for example some properties of star formation, as the initial mass function (IMF), could change systematically with galaxy mass (\\cite{koppen07}). All these effects have a deep impact on galaxy formation, and the knowledge of their relative contributions is of crucial importance. Different models have been built to reproduce the shape of the mass-metallicity relation in the local universe, and different assumptions produce divergent predictions at high redshifts (z$>$2). To explore this issue several groups have observed the mass-metallicity relation in the distant universe, around z=0.7 (\\cite{savaglio05}) and z=2.2 (\\cite{erb06}). They have found a clear evolution with cosmic time, with metallicity for a given stellar mass decreasing with increasing redshift. For several reasons, it is very interesting to explore even higher redshifts. The redshift range at z$\\sim$3--4 is particularly interesting: it is before the peak of the cosmic star formation density (see, for example, \\cite{mannucci07}), only a small fraction ($~$15\\%, \\cite{pozzetti07}) of the total stellar mass have already been created, the number of mergers among the galaxies is much larger than at later times (\\cite{conselice07}). As a consequence, the prediction of the different models tend to diverge above z=3, and it is important to sample this redshift range observationally. The observations are really challenging because of the faintness of the targets and the precision required to obtain a reliable metallicity. Nevertheless, the new integral-field unit (IFU) instruments on 8-m class telescopes are sensitive enough to allow for the project. ", "conclusions": "" }, "0808/0808.4126_arXiv.txt": { "abstract": "The carrier of the 2175$\\Angstrom$ interstellar extinction feature remains unidentified since its first detection over 40 years ago. In recent years carbon buckyonions have been proposed as a carrier of this feature, based on the close similarity between the electronic transition spectra of buckyonions and the 2175$\\Angstrom$ interstellar feature. We examine this hypothesis by modeling the interstellar extinction with buckyonions as a dust component. It is found that dust models containing buckyonions (in addition to amorphous silicates, polycyclic aromatic hydrocarbon molecules, graphite) can closely reproduce the observed interstellar extinction curve. To further test this hypothesis, we call for experimental measurements and/or theoretical calculations of the infrared vibrational spectra of hydrogenated buckyonions. By comparing the infrared emission spectra predicted for buckyonions vibrationally excited by the interstellar radiation with the observed emission spectra of the diffuse interstellar medium, we will be able to derive (or place an upper limit on) the abundance of interstellar buckyonions. ", "introduction": "In the interstellar medium (ISM), the strongest spectroscopic extinction feature is the 2175$\\Angstrom$ bump which is characterized with a stable peak wavelength and a width variable from one sightline to another (Fitzpatrick \\& Massa 2007). Since Stecher (1965) first detected this ultraviolet (UV) extinction feature through rocket observations, the origin of this feature and the nature of its carrier(s) are still an enigma. Many candidate materials, including graphite, amorphous carbon, graphitized (dehydrogenated) hydrogenated amorphous carbon, nano-sized hydrogenated amorphous carbon, quenched carbonaceous composite, coals, polycyclic aromatic hydrocarbon (PAH), and OH$^{-}$ ion in low-coordination sites on or within silicate grains have been proposed, while no single one is generally accepted (see Li \\& Greenberg 2003 for a review). Recently, Chhowalla et al.\\ (2003) measured the UV-visible photoabsorption spectra of carbon buckyonions (BOs)\\footnote{% Early pioneering experimental studies of the UV absorption properties of BOs and their association with the 2175$\\Angstrom$ interstellar extinction feature include those of de Heer \\& Ugarte (1993), Ugarte (1995), and Wada et al.\\ (1999). } composed of spherical concentric fullerene shells (so far the largest BOs produced in laboratory have $N\\sim 100$ shells; Iglesias-Groth et al.\\ 2003). They found that the plasmon-like feature of BOs due to a collective excitation of the $\\pi$ electrons closely fits the 2175$\\Angstrom$ interstellar extinction feature. More recently, Ruiz et al.\\ (2005) theoretically simulated the photoabsorption spectra of BOs.\\footnote{% Early pioneering theoretical studies of the UV absorption properties of BOs and their association with the 2175$\\Angstrom$ interstellar extinction feature include those of Wright (1988), Henrard et al.\\ (1993, 1997) and Lucas et al.\\ (1994). } They found that the calculated absorption spectra of BOs are in close agreement with the experimental data of Chhowalla et al.\\ (2003) and the observed 2175$\\Angstrom$ interstellar feature. That the $\\pi$-plasmon absorption band of BOs exhibits a stable peak position at $\\simali$5.70\\,eV ($\\lambda\\approx 2175\\Angstrom$, $\\lambda^{-1}\\approx 4.6\\mum^{-1}$) and a variable bandwidth (Chhowalla et al.\\ 2003, Ruiz et al.\\ 2005) suggests that BOs may be a promising candidate material for the 2175$\\Angstrom$ interstellar extinction feature. Indeed, it is shown by Chhowalla et al.\\ (2003) and Ruiz et al.\\ (2005) that the photoabsorption spectra of BOs {\\it alone} very well reproduce the entire interstellar extinction curve at $\\lambda^{-1}$\\,$\\simali$3.2--7.3$\\mum^{-1}$. This requires $\\simali$190\\,ppm (parts per million) C/H to be locked up in BOs (Ruiz et al.\\ 2005). However, it is well recognized that in the ISM, in addition to the 2175$\\Angstrom$ extinction carrier, there must exist other dust components as well -- there must be a population of amorphous silicate dust, as indicated by the strong, ubiquitous 9.7 and 18$\\mum$ interstellar absorption features; there must be a population of aromatic hydrocarbon dust (presumably PAH molecules), as indicated by the distinctive set of ``unidentified'' infrared (UIR) emission bands at 3.3, 6.2, 7.7, 8.6, and 11.3$\\mum$ ubiquitously seen in the ISM; there must also exist a population of aliphatic hydrocarbon dust, as indicated by the 3.4$\\mum$ C--H absorption feature which is also ubiquitously seen in the diffuse ISM of the Milky Way and external galaxies (see Li 2004). Although buckyonions are able to closely reproduce the 2175$\\Angstrom$ extinction feature, it is not clear if dust models with amorphous silicates, PAHs, and other carbon dust species (e.g. amorphous carbon, hydrogenated amorphous carbon, organic refractory, and graphite) incorporated (in addition to BOs) are still capable of fitting the 2175$\\Angstrom$ extinction feature. One may intuitively expect that the almost perfect fit to the $\\lambda^{-1}$\\,$\\simali$3.2--7.3$\\mum^{-1}$ interstellar extinction would easily be distorted by the addition of other dust components. It is the purpose of this {\\it Letter} to examine this issue. To this end, we consider the extinction of dust models consisting of multi-modal grain populations with BOs as an interstellar grain component. \\begin{figure} \\begin{center} \\includegraphics{f1.eps} \\end{center} \\caption{\\label{fig:mrn_solar} Comparison of the interstellar extinction (dot-dashed line) with the model extinction (solid line with oscillatory features) obtained by summing up the contributions of buckyonions (C/H\\,=\\,115\\,ppm), PAHs (C/H\\,=\\,60\\,ppm), graphite (C/H\\,=\\,35\\,ppm), and amorphous silicate (Si/H\\,=\\,35\\,ppm). We assume a MRN-type size distribution for the silicate and graphite grains. } \\end{figure} ", "conclusions": "It is encouraging that the dust models consisting of multi-modal grain populations (including BOs) fit the interstellar extinction curve (including the 2175$\\Angstrom$ extinction feature) very well while satisfying the interstellar abundance constraints. The fit to the 2175$\\Angstrom$ feature is mainly affected by the silicate dust component. The smaller amount of silicate dust is included in the model, the better is the fit. We would achieve a closer match to the 2175$\\Angstrom$ extinction if we take the interstellar Si abundance to be subsolar (Snow \\& Witt 1996) like that of B stars $\\sidust=18\\ppm$ (Sofia \\& Meyer 2001) while having C/H\\,=\\,140\\,ppm in BOs and C/H\\,=\\,10\\,ppm in graphite.\\footnote{% If both C and Si are subsolar (say, 2/3 of that of solar, see Snow \\& Witt 1996), we will not be able to fit the interstellar extinction (also see Li 2005). } BOs are a promising candidate material for the 2175$\\Angstrom$ interstellar extinction feature: (1) their $\\pi$-plasmon absorption profile closely resembles the 2175$\\Angstrom$ extinction feature and is stable in peak wavelength position while varies in width;\\footnote{% Ruiz et al.\\ (2005)'s theoretically simulated photoabsorption spectra of BOs showed that the width of the 2175$\\Angstrom$ feature $\\gammabump$ varies with size for small BOs, however, the width becomes independent of size when the number of shells $N\\simgt 6$ ($\\gammabump \\approx 1.03\\mum^{-1}$; see their Fig.\\,4). For BOs to explain the 2175$\\Angstrom$ interstellar features broader than $\\gammabump = 1.03\\mum^{-1}$ (some sightlines have $\\gammabump$ up to $\\simali$1.2$\\mum^{-1}$), clustering of individual BOs (de Heer \\& Ugarte 1993, Rouleau et al.\\ 1997), imperfect growth of BOs (e.g. mixing with amorphous carbon impurities; see Chhowalla et al.\\ 2003), and coating BOs with a layer of PAHs (Mathis 1994) may play an important role. } (2) the almost perfect fit to the interstellar feature is not significantly distorted by the inclusion of other dust components (e.g. silicate, PAHs, graphite) which are required to account for other interstellar phenomena (e.g. the 9.7, 18$\\mum$ absorption features, the ``UIR'' bands); and (3) BOs are highly stable molecules that they can survive under intense UV radiation and are highly resistant to destruction by collisions. BOs can be formed by electron beam irradiation of carbon soot (Ugarte 1992) and by heat treatment (annealing at temperatures $T\\simgt 700\\,^{\\circ}$C) and electron beam irradiation of nanodiamond (Kuznetsov et al.\\ 1994, Tomita et al.\\ 2002). The generation of BOs by annealing nanodiamonds is astrophysically relevant. Presolar nanodiamonds are identified in primitive carbonaceous meteorites based on their isotopic anomalies (Lewis et al.\\ 1987). The possible existence of nanodiamonds in dense clouds and circumstellar dust disks or envelopes have also been suggested (Allamandola et al.\\ 1992, Guillois, Ledoux, \\& Reynaud 1999, van Kerckhoven, Tielens, \\& Waelkens 2002). The generation of BOs through annealing nanodiamonds can occur in the ISM where nanodiamonds are stochastically heated to temperatures as high as $\\simali$1000$\\K$ by the interstellar radiation field (Jones \\& d'Hendecourt 2000). Indeed, BOs have been found in meteorites with anomalous isotopic compositions (Smith \\& Buseck 1981, Bernatowicz et al.\\ 1996, Harris et al.\\ 2000). Recently, BOs have also been invoked to explain other interstellar phenomena: the mysterious diffuse interstellar bands (DIBs) resulting from the electronic transitions of BOs (Iglesias-Groth 2004, 2007),\\footnote{% To date, $>$300 DIBs have been detected in the Galactic and extragalactic ISM. These mysterious optical-to-near-IR absorption lines (broader than the narrow lines from gas atoms, ions, and small molecules) still remain unidentified. Fullerenes are known to have strong structural resonances at this wavelength range (Dresselhaus et al.\\ 1996, Iglesias-Groth 2004). Iglesias-Groth (2007) argued that BOs may be responsible for the strongest optical DIB at 4430$\\Angstrom$ and possibly also the 6177$\\Angstrom$ and 6284$\\Angstrom$ DIBs. } the 10--100\\,GHz Galactic anomalous microwave emission resulting from the electric dipole emission of rotationally excited BOs (Iglesias-Groth 2005, 2006), and the broad feature around 100$\\mum$ of the diffuse emission from two active star-forming regions (the Carina Nebula and Sharpless 171) resulting from the small particle surface resonance (Onaka \\& Okada 2003). Experimentally, all BOs are found to belong to the C$_{60N^2}$ icosahedral family with $N=1,2,...$ (i.e. their shells are consecutive elements of the icosahedral C$_{60N^2}$ fullerene family, with C$_{60}$ being the smallest shell\\footnote{% Kroto et al.\\ (1985) first proposed that C$_{60}$ could be present in the ISM with a considerable quantity. This molecule and its related species were later proposed as the carriers of the 2175$\\Angstrom$ extinction hump, the DIBs, the ``UIR'' bands, and the extended red emission (see Webster 1991, 1992, 1993a,b). C$_{60}$ and C$_{70}$ are unlikely a major contributor to the 2175$\\Angstrom$ extinction feature since they have a characteristic doublet absorption in this wavelength region. Foing \\& Ehrenfreund (1994) attributed the two DIBs at 9577$\\Angstrom$ and 9632$\\Angstrom$ to C$_{60}^{+}$. However, attempts to search for these molecules in the UV and IR were unsuccessful (Snow \\& Seab 1989; Somerville \\& Bellis 1989; Moutou et al.\\ 1999; Herbig 2000). These molecules are now estimated to consume at most $<$0.7$\\ppm$ carbon (Moutou et al.\\ 1999). Therefore, C$_{60}$ is at most a minor component of the interstellar dust family. } and the intershell distance being very close to the interplanar separation in graphite, often taken to be 3.55$\\Angstrom$, the C$_{60}$ radius; Yoshida \\& 1993, Ruiz et al.\\ 2004). So far, all the experimentally generated and analyzed BOs are in the nanometer size range ($\\simali$3--15\\,nm in diameter; see de Heer \\& Ugarte 1993, Cabioc'h et al.\\ 1997, Chhowalla et al.\\ 2003). The fullerenes produced in laboratory are usually accompanied by tubular and other non-spherical graphite-like structures of which the theoretical UV spectra exhibit some similarity to that of BOs. In the ISM, tubular graphite and PAHs can be catalytically formed at $T\\simgt 1000\\K$ on Fe nanoparticles and are probably the carrier of some of the DIBs (Zhou et al.\\ 2006). They may also contribute to the 2175$\\Angstrom$ interstellar extinction feature. Finally, we should note that a powerful test of the BOs hypothesis would be in the IR. Because of their small sizes (and therefore small heat capacities), in the ISM, BOs will be stochastically heated by single UV photons (Draine \\& Li 2001). With their surface shell hydrogenated, they will emit in the near- to mid-IR through their characteristic C--H stretching and bending bands, and C--C stretching bands. The detection (or non-detection) of these bands will allow us to derive (or place an upper limit on) the abundance of BOs. Note that the BO models require an appreciable amount of C/H to be in BOs ($>110\\ppm$), nearly twice as much as PAHs. Unfortunately, little is known about the positions and strengths of these vibrational bands. We call for urgent experimental measurements and/or theoretical calculations of the IR vibrational spectra of hydrogenated BOs. Future far-IR/submm space missions (e.g. Herschel) will also be useful for studying their far-IR vibrational spectra (see Iglesias-Groth \\& Bret\\'on 2000). Moreover, it would be interesting to see if the silicate-graphite-PAH-BO multi-component dust model is able to reproduce the $\\simali$2--3000$\\mum$ overall IR emission of the Galactic ISM." }, "0808/0808.0914_arXiv.txt": { "abstract": "We present Space Telescope Imaging Spectrograph observations of 14 nearby low-luminosity active galactic nuclei, including 13 LINERs and 1 Seyfert, taken at multiple parallel slit positions centered on the galaxy nuclei and covering the H$\\alpha$ spectral region. For each galaxy, we measure the emission-line velocities, line widths, and strengths, to map out the inner narrow-line region structure, typically within $\\sim$100 pc from the galaxy nucleus. There is a wide diversity among the velocity fields: in a few galaxies the gas is clearly in disk-like rotation, while in other galaxies the gas kinematics appear chaotic or are dominated by radial flows with multiple velocity components. In most objects, the emission-line surface brightness distribution is very centrally peaked. The [\\ion{S}{2}] line ratio indicates a radial stratification in gas density, with a sharp increase within the inner 10--20 pc, in the majority of the Type 1 (broad-lined) objects. The electron-density gradients of the Type 1 objects exhibit a similar shape that is well fit by a power law of the form $\\emph{n}_{\\mathrm{e}} = \\emph{n}_0(r/1~ \\mathrm{pc})^{\\alpha}$, where $\\alpha = -0.60 \\pm{0.13}$. We examine how the [\\ion{N}{2}] $\\lambda 6583$ line width varies as a function of aperture size over a range of spatial scales, extending from scales comparable to the black hole's sphere of influence to scales dominated by the host galaxy's bulge. For most galaxies in the sample, we find that the emission-line velocity dispersion is largest within the black hole's gravitational sphere of influence, and decreases with increasing aperture size toward values similar to the bulge stellar velocity dispersion measured within ground-based apertures. We construct models of gas disks in circular rotation and show that this behavior can be consistent with virial motion, although for some combinations of disk parameters we show that the line width can increase as a function of aperture size, as observed in NGC 3245. Future dynamical modeling in order to determine black hole masses for a few objects in this sample may be worthwhile, although disorganized motion will limit the accuracy of the mass measurements. ", "introduction": "\\label{sec:intro} With its $\\sim$ 0\\farcs1 resolution, the Space Telescope Imaging Spectrograph (STIS) aboard the \\emph{Hubble Space Telescope (HST)} has provided the exceptional ability to spatially resolve gas kinematics within the gravitational sphere of influence of the putative central black hole in nearby galaxies. Resolving the gas kinematics very close to the black hole allows for the determination of black hole mass in a relatively straightforward manner, provided that the gas is in Keplerian rotation in a thin, disk-like structure. STIS has thus had a major impact on the field of supermassive black hole detection in galactic nuclei, and data from STIS have contributed greatly toward establishing the correlations between black hole mass and host-galaxy properties, such as those with the stellar velocity dispersion \\citep{Ferrarese_2000,Gebhardt_2000,Tremaine_2002} and bulge luminosity \\citep{Kormendy_Gebhardt_2001}. More recently, ground-based integral field unit (IFU) observations have provided two-dimensional (2D) kinematic information on the ionized gas in galaxy centers \\citep{Sarzi_2006, Dumas_2007}. Additionally, ground-based IFU observations with the assistance of adaptive optics have produced gas dynamical black hole mass measurements \\citep{Hicks_Malkan_2007, Neumayer_2007}. Without the use of adaptive optics, however, ground-based observations are usually unable to resolve scales comparable to the black hole radius of influence in nearby galaxies. Although STIS is a long-slit instrument, 2D information similar to IFU data can be obtained with STIS by using multiple slit positions. Even with the past decade of high-resolution observations, only about 30 dynamical black hole mass measurements have been made \\citep{Ferrarese_Ford_2005}. Consequently, searching for regular gas velocity fields dominated by gravitational motion is essential for extending the number of black hole mass measurements and further establishing the connections between black holes and host galaxies. While searching for regular gas velocity fields is clearly important, it appears, however, that they are fairly rare. Instead, complicated and chaotic emission-line velocity fields are more commonly found near the centers of galaxies \\citep{Ho_2002, Atkinson_2005}. Although only upper limits on the black hole mass can be estimated using gas dynamical modeling in these situations \\citep{Sarzi_2002}, subarcsecond-resolution observations can still be useful for studying the dynamical state of the gas and searching for radial motions that might be related to active galactic nucleus (AGN) fueling or outflows. Modeling AGN outflows on subarcsecond scales \\citep{Capetti_1996, Crenshaw_Kraemer_2000} is important for furthering our understanding of the connection between AGN feedback and host-galaxy properties. Of particular interest are the connection to star formation, the growth of the supermassive black hole, and how gas may be driven out from the host galaxy as predicted in some feedback scenarios \\citep{Springel_2005, Hopkins_2006, Sijacki_2007}. Moreover, the narrow-line region (NLR) velocity dispersion ($\\sigma_{g}$), often measured from the width of the [\\ion{O}{3}] $\\lambda 5007$ line, may provide a reasonable estimate of the host-galaxy stellar velocity dispersion ($\\sigma_\\star$), albeit with substantial scatter \\citep{Nelson_Whittle_1996, Greene_Ho_2005}. Detailed NLR studies are essential in understanding the origin of the scatter in the $\\sigma_{g}$--$\\sigma_\\star$ relationship. This relationship can be useful for distant and luminous AGNs where measurement of the stellar velocity dispersion is not feasible \\citep{Salviander_2007, Shields_2003}. Using the narrow emission-line widths as a proxy for stellar velocity dispersion is only meaningful if the NLR gas is dominated by virial motion in the host-galaxy bulge and not by the gravitational influence of the black hole or nongravitational motions. Therefore, outflows and similar nongravitational motions, which can disrupt viral motion, are one potential cause of the scatter in the $\\sigma_{g}$--$\\sigma_\\star$ relationship. While nongravitational motions are clearly important in setting the narrow-line widths in AGNs with strong outflows, there can be an important contribution to the line widths from nongravitational motion even in objects that are not outflow-dominated. From \\emph{HST} STIS observations, \\cite{VerdoesKleijn_2006} found that circumnuclear gas in Fanaroff-Riley 1 radio galaxies typically has an intrinsic velocity dispersion in excess of that expected from purely gravitational motion in the host-galaxy potential. They concluded that this excess line width is driven by the injection of energy by the AGN, although there was no obvious relationship between the AGN radiative luminosity and the magnitude of the nongravitational line dispersion. Other origins for the scatter were studied by \\cite{Greene_Ho_2005} using a large sample of Sloan Digital Sky Survey (SDSS) type 2 AGNs. They looked for correlations of nuclear properties (radio power, AGN luminosity, Eddington ratio), as well as global properties (host-galaxy morphology, local environment, star-formation rate), with the excess line width, defined to be the difference between the dispersion of the [\\ion{O}{3}] $\\lambda 5007$ line and the stellar velocity dispersion. They found a strong correlation between the excess line width and Eddington ratio, where larger excess line widths, indicative of outflowing gas, were seen in the higher Eddington ratio objects. Another potential explanation for the scatter in the $\\sigma_{g}$--$\\sigma_\\star$ relationship is the effect of observational aperture size. \\cite{Rice_2006} investigated the relationship between line width and aperture size in a sample of nearby Seyferts using single-slit STIS observations. They found a wide range of behavior in how the widths of the [\\ion{O}{3}] $\\lambda 5007$ and [\\ion{S}{2}] $\\lambda\\lambda$6716, 6731 lines varied as a function of radius, but the most common trends were either a roughly constant width or an increase in width with increasing aperture size. Their analysis showed that the line widths measured within the largest STIS aperture were systematically smaller by 10\\%--20\\% than both the stellar velocity dispersion and the line widths measured within a ground-based sized aperture. They concluded that the smaller line width measured within the randomly oriented STIS slit is the result of sampling a smaller portion of the velocity field than the larger ground-based aperture. Beyond the $\\sigma_{g}$--$\\sigma_\\star$ relationship, we can address other open questions concerning low-luminosity AGNs. \\cite{Laor_2003} pointed out that while the NLR in luminous AGNs occurs on scales of tens of parsecs, where its dynamics are dominated by the bulge, the NLR in low-luminosity AGNs is more compact, and its dynamics could be dominated by the black hole. It is therefore interesting to ask whether the narrow-line widths in low-luminosity AGNs are set by the black hole or by the bulge potential. The NLR emission lines also provide information on ionization mechanisms. Spectra of low-ionization nuclear emission-line regions \\citep[LINERs;][]{Heckman_1980} can be reproduced with photoionization models involving a nonstellar continuum and low ionization parameters \\citep{Ferland_Netzer_1983, Halpern_Steiner_1983}, but high electron densities, as well as a large range of densities, are needed in order to reproduce the strength of the [\\ion{O}{3}] $\\lambda 4363$ line \\citep{Filippenko_Halpern_1984, Filippenko_1985, Ho_1993}. \\cite{Filippenko_Halpern_1984} demonstrated that such conditions exist in the LINER NGC 7213, and that there is a correlation between line width and the critical density for collisional deexcitation. A similar relationship between line width and ionization potential was also previously seen in Seyferts \\citep{Pelat_1981}. This implied that the NLR is not a homogeneous environment, but must have radial gradients in velocity and density. In ground-based spectra these spatial gradients were generally unresolved, but using STIS it is possible to directly resolve the density gradients in the inner NLR \\citep{Barth_2001a, Shields_2007}. It is therefore of interest to determine what fraction of LINERs show spatially resolved density gradients, and whether the type 2 LINERs, where a central photoionizing source is not always seen, exhibit density gradients. In this paper, we present multislit \\emph{HST} STIS observations of 14 low-luminosity AGNs, including 13 LINERs and 1 Seyfert. We use the high spatial resolution of STIS to map out the 2D kinematic structure of the NLR within the inner $\\sim 100$ pc, and search for velocity fields that show signs of regular rotation for which gas dynamical modeling may be performed. The subarcsecond resolution STIS data also allow us to detect electron-density gradients through the [\\ion{S}{2}] $\\lambda 6716/\\lambda 6731$ line ratio diagnostic. Finally, we examine how the [\\ion{N}{2}] line width varies as a function of aperture size over spatial scales ranging from those comparable to the black hole sphere of influence to scales dominated by the bulge. ", "conclusions": "\\label{sec:conclusions} We have used STIS data to measure detailed velocity, velocity dispersion, flux, and line-intensity ratio information within $\\sim 100$ pc from the nucleus of 14 nearby galaxies. All of the galaxies harbor low-luminosity AGNs, 13 of which are classified as LINERs and one of which is a Seyfert. We see a large diversity in the velocity fields of the 14 galaxies, where simple disk rotation is not a common state. Only a few galaxies have definite disk rotation, while some others have possible disk rotation but the emission-line S/N is too low to clearly determine the kinematic state of the gas. Some velocity fields are dominated by irregular motions, some show an overall rotation but with large random components, and others clearly exhibit outflows. Dynamical modeling of NGC 2911 and NGC 4594 to determine black hole masses may be possible and worthwhile, although chaotic and disorganized motion will limit the accuracy of mass measurements. Through the [\\ion{S}{2}] $\\lambda 6716$/[\\ion{S}{2}] $\\lambda 6731$ diagnostic, we were able to detect significant electron-density gradients in four (NGC 1052, NGC 3998, NGC 4278, and NGC 4579) of the five LINERs having a sufficient S/N. Moreover, the electron-density gradients in these objects share a similar shape that can be described by a power law with a slope of $-0.60 \\pm {0.13}$. A similar increase near the nucleus was seen in the [\\ion{O}{1}]/[\\ion{S}{2}] line ratio for the first three of these galaxies. These results demonstrate that density gradients within the inner $\\sim 20$ pc are common in type 1 LINERs. This is consistent with expectations from photoionization models of LINERs and from ground-based detections of correlations between line width and \\emph{n}$_{\\mathrm{crit}}$ in LINERs. We find that the density gradients persist in NGC 1052 and over a portion of the NLR in NGC 3227, even in the presence of strong outflows that clearly have a large impact on the NLR gas as seen through the disorganized velocity fields with multiple velocity components. We also see evidence for correlations between line width and \\emph{n}$_{\\mathrm{crit}}$ in most LINERs even on the 0\\farcs1 scales probed by STIS. Measurements of the STIS spectra additionally provided a means to study the [\\ion{N}{2}] $\\lambda 6583$ narrow-line width variation with spectrograph aperture size. For most galaxies in our sample, the emission-line velocity dispersion peaks within the black hole sphere of influence, then decreases in larger apertures, approaching the bulge stellar velocity dispersion. This trend is consistent with, but does not uniquely imply, virial motion. The increasing line width with aperture size seen in NGC 1052 and NGC 3227 is due to the outflows that dominate the NLR kinematics. In NGC 3245, the increase is the result of the combination of the galaxy's fairly flat emission-line surface brightness profile, the galaxy's massive bulge, and the highly inclined gaseous disk. NGC 3245 demonstrates that even gas in pure disk rotation can exhibit an increasing line width as a function of aperture size. While work by \\cite{Boroson_2005} and \\cite{Greene_Ho_2005} provide evidence for a connection between the Eddington ratio and kinematic disturbance in the NLR, which may reflect an underlying trend of outflow-dominated NLRs being more predominant at high AGN luminosities, our sample of galaxies is too small and does not span a sufficiently wide luminosity range to arrive at such a conclusion. Since such a relationship would be helpful in understanding AGN feedback and the effects on host-galaxy properties, further investigation, ideally with IFU data on a much larger sample of AGNs spanning a wide range of luminosities, is needed." }, "0808/0808.2250_arXiv.txt": { "abstract": "Pre-main-sequence stars are observed to be surrounded by both accretion flows and some kind of wind or jet-like outflow. Recent work by Matt and Pudritz has suggested that if classical T Tauri stars exhibit stellar winds with mass loss rates about 0.1 times their accretion rates, the wind can carry away enough angular momentum to keep the stars from being spun up unrealistically by accretion. This paper presents a preliminary set of theoretical models of accretion-driven winds from the polar regions of T Tauri stars. These models are based on recently published self-consistent simulations of the Sun's coronal heating and wind acceleration. In addition to the convection-driven MHD turbulence (which dominates in the solar case), we add another source of wave energy at the photosphere that is driven by the impact of plasma in neighboring flux tubes undergoing magnetospheric accretion. This added energy, determined quantitatively from the far-field theory of MHD wave generation, is sufficient to produce T Tauri-like mass loss rates of at least 0.01 times the accretion rate. While still about an order of magnitude below the level required for efficient angular momentum removal, these are the first self-consistent models of T Tauri winds that agree reasonably well with a range of observational mass loss constraints. The youngest modeled stellar winds are supported by Alfv\\'{e}n wave pressure, they have low temperatures (``extended chromospheres''), and they are likely to be unstable to the formation of counterpropagating shocks and clumps far from the star. ", "introduction": "Our current state of knowledge about how stars and planets are formed comes from an intertwined web of observations (spanning the electromagnetic spectrum) and theoretical work. The early stages of low-mass star formation comprise a wide array of inferred physical processes, including disk accretion, various kinds of outflow, and magnetohydrodynamic (MHD) activity on time scales ranging from hours to millennia (see, e.g., Lada 1985; Bertout 1989; Appenzeller \\& Mundt 1989; Hartmann 2000; K\\\"{o}nigl \\& Pudritz 2000; McKee \\& Ostriker 2007; Shu et al.\\ 2007). A key recurring issue is that there is a great deal of {\\em mutual interaction and feedback} between the star and its circumstellar environment. This interaction can determine how rapidly the star rotates, how active the star appears from radio to X-ray wavelengths, and how much mass and energy the star releases into its interplanetary medium. A key example of circumstellar feedback is the magnetospheric accretion paradigm for classical T Tauri stars (Lynden-Bell \\& Pringle 1974; Uchida \\& Shibata 1984; Camenzind 1990; K\\\"{o}nigl 1991). Because of strong ($\\sim$1 kG) stellar magnetic fields, the evolving equatorial accretion disk does not penetrate to the stellar surface, but instead is stopped by the stellar magnetosphere. Accretion is thought to proceed via ballistic infall along magnetic flux tubes threading the inner disk, leading to shocks and hot spots on the surface. The primordial accretion disk is dissipated gradually as the star enters the weak-lined T Tauri star phase, with a likely transition to a protoplanetary dust/debris disk. Throughout these stages, solar-mass stars are inferred to exhibit some kind of wind or jet-like outflow. There are several possible explanations of how and where the outflows arise, including extended disk winds, X-winds, impulsive (plasmoid-like) ejections, and ``true'' stellar winds (e.g., Paatz \\& Camenzind 1996; Calvet 1997; K\\\"{o}nigl \\& Pudritz 2000; Dupree et al.\\ 2005; Edwards et al.\\ 2006; Ferreira et al.\\ 2006; G\\'{o}mez de Castro \\& Verdugo 2007; Cai et al.\\ 2008). Whatever their origin, the outflows produce observational diagnostics that indicate mass loss rates exceeding the Sun's present mass loss rate by factors of $10^3$ to $10^6$. It is of some interest to evaluate how much of the observed outflow can be explained solely with {\\em stellar} winds, since these flows are locked to the star and thus are capable of removing angular momentum from the system. Recent work by Matt \\& Pudritz (2005, 2007, 2008) has suggested that if there is a stellar wind with a sustained mass loss rate about 10\\% of the accretion rate, the wind can carry away enough angular momentum to keep T Tauri stars from being spun up unrealistically by the accretion. Despite many years of study, the dominant physical processes that accelerate winds from cool stars have not yet been identified conclusively. For many stars, the large-scale energetics of the system---i.e., the luminosity and the gravitational potential---seem to determine the overall magnitude of the mass loss (Reimers 1975; 1977; Schr\\\"{o}der \\& Cuntz 2005). Indeed, for the most luminous cool stars, radiation pressure seems to provide a direct causal link between the luminosity $L_{\\ast}$ and the mass loss rate $\\dot{M}_{\\rm wind}$ (e.g., Gail \\& Sedlmayr 1987; H\\\"{o}fner 2005). However, for young solar-mass stars, other mediating processes (such as coronal heating, waves, or time-variable magnetic ejections) are more likely to connect the properties of the stellar interior to the surrounding outflowing wind. For example, magnetohydrodynamic (MHD) waves have been studied for several decades as a likely way for energy to be transferred from late-type stars to their winds (Hartmann \\& MacGregor 1980; DeCampli 1981; Airapetian et al.\\ 2000; Falceta-Gon\\c{c}alves et al.\\ 2006; Suzuki 2007). In parallel with the above work in improving our understanding of stellar winds, there has been a great deal of progress toward identifying and characterizing the processes that produce the {\\em Sun's} corona and wind. It seems increasingly clear that closed magnetic loops in the low solar corona are heated by small-scale, intermittent magnetic reconnection that is driven by the continual stressing of their footpoints by convective motions (e.g., Aschwanden 2006; Klimchuk 2006). The open field lines that connect the Sun to interplanetary space, though, appear to be energized by the dissipation of waves and turbulent motions that originate at the stellar surface (Tu \\& Marsch 1995; Cranmer 2002; Suzuki 2006; Kohl et al.\\ 2006). Parker's (1958) classic paradigm of solar wind acceleration via gas pressure in a hot ($T \\sim 10^{6}$ K) corona still seems to be the dominant mechanism, though waves and turbulence have an impact as well. A recent self-consistent model of turbulence-driven coronal heating and solar wind acceleration has succeeded in reproducing a wide range of observations {\\em with no ad-hoc free parameters} (Cranmer et al.\\ 2007). This progress on the solar front is a fruitful jumping-off point for a better understanding of the basic physics of winds and accretion in young stars. The remainder of this paper is organized as follows. {\\S}~2 presents an overview of the general scenario of accretion-driven MHD waves that is proposed here to be important for driving T Tauri mass loss. In {\\S}~3 the detailed properties of an evolving solar-mass star are presented, including the fundamental stellar parameters, the accretion rate and disk geometry, and the properties of the clumped gas in the magnetospheric accretion streams. These clumped streams impact the stellar surface and create MHD waves that propagate horizontally to the launching points of stellar wind streams. {\\S}~4 describes how self-consistent models of these wind regions are implemented, and {\\S}~5 gives the results. Finally, {\\S}~6 contains a summary of the major results of this paper and a discussion of the implications these results may have on our wider understanding of low-mass star formation. ", "conclusions": "The primary aim of this paper has been to show how accretion-driven waves on the surfaces of T Tauri stars may help contribute to the strong rates of atmospheric heating and large mass loss rates inferred for these stars. The ZEPHYR code, which was originally developed to model the solar corona and solar wind (Cranmer et al.\\ 2007), has been applied to the T Tauri stellar wind problem. A key aspect of the models presented above is that the only true free parameters are: (1) the properties of MHD waves injected at the photospheric lower boundary, and (2) the background magnetic geometry. Everything else (e.g., the radial dependence of the rates of chromospheric and coronal heating, the temperature structure of the atmosphere, and the wind speeds and mass fluxes) emerges naturally from the modeling process. For solar-mass T Tauri stars, time-steady stellar winds were found to be supportable for all ages older than about 0.45 Myr, with accretion rates less than $7 \\times 10^{-8}$ $M_{\\odot}$ yr$^{-1}$ driving mass loss rates less than $4 \\times 10^{-10}$ $M_{\\odot}$ yr$^{-1}$. Still younger T Tauri stars (i.e., ages between about 13 kyr and 0.45 Myr) may exhibit time-variable winds with mass loss rates extending up several more orders of magnitude to $\\sim 2 \\times 10^{-8}$ $M_{\\odot}$ yr$^{-1}$. The transition between time-steady and variable winds occurs when the critical point of the flow migrates far enough past the Alfv\\'{e}n point (at which the wind speed equals the Alfv\\'{e}n speed) such that the Alfv\\'{e}n wave amplitude begins to decline rapidly with increasing radius. When this happens, the outward wave-pressure acceleration is quickly ``choked off;'' i.e., parcels of gas that make it past the critical point cannot be accelerated to infinity, and stochastic collisions between upflowing and downflowing parcels must begin to occur. The maximum wind efficiency ratio $\\dot{M}_{\\rm wind}/\\dot{M}_{\\rm acc}$ for the T Tauri models computed here was approximately 1.4\\%, computed for ages of order 0.1 Myr. This is somewhat smaller than the values of order 10\\% required by Matt \\& Pudritz (2005, 2007, 2008) to remove enough angular momentum from the young solar system to match present-day conditions. It is also well below the observational ratios derived by, e.g., Hartigan et al.\\ (1995, 2004), which can reach up to 20\\% for T Tauri stars between 1 and 10 Myr old (see Fig.~12). These higher ratios, though, may be the product of both stellar winds and disk winds (possibly even dominated by the disk wind component). Additionally, it is possible that future observational analysis will result in these empirical ratios being revised {\\em upward} with more accurate (lower) values of $\\dot{M}_{\\rm acc}$ (S.\\ Edwards 2008, private communication). The accretion-driven solutions for $\\dot{M}_{\\rm wind}$ depend crucially on the properties of the waves in the polar regions. It is important to note that the calculation of MHD wave properties was based on several assumptions that should be examined in more detail. The relatively low MHD wave efficiency used in equation (\\ref{eq:EA}) is an approximation based on the limiting case of waves being far from the impact site. A more realistic model would have to contain additional information about both the nonlinearities of the waves themselves and the vertical atmospheric structure through which the waves propagate. It seems likely that a better treatment of the wave generation would lead to larger wave energies at the poles. On the other hand, the assumption that the waves do not damp significantly between their generation point and the polar wind regions may have led to an assumed wave energy that is too high. It is unclear how strong the waves will be in a model that takes account of all of the above effects. In any case, the ZEPHYR results presented here are the first self-consistent models of T Tauri stellar winds that produce wind efficiency ratios that even get into the right ``ballpark'' of the angular momentum requirements. Additional improvements in the models are needed to make further progress. For example, the effects of stellar rotation should be included, both in the explicit wind dynamics (e.g., ``magneto-centrifugal driving,'' as recently applied by Holzwarth \\& Jardine 2007) and in the modified subsurface convective activity that is likely to affect photospheric amplitudes of waves and turbulence. Young and rapidly rotating stars are likely to have qualitatively different convective motions than are evident in standard (nonrotating, mixing-length) models. It is uncertain, though, whether rapid rotation gives rise to larger (K\\\"{a}pyl\\\"{a} et al.\\ 2007; Brown et al. 2007; Ballot et al. 2007) or smaller (Chan 2007) convection eddy velocities at the latitudes of interest for T Tauri stellar winds. In any case, rapid rotation can also increase the buoyancy of subsurface magnetic flux elements, leading to a higher rate of flux emergence (Holzwarth 2007). Also, the lower gravities of T Tauri stars may give rise to a larger fraction of the convective velocity reaching the surface as wave energy (e.g., Renzini et al.\\ 1977), or the convection may even penetrate directly into the photosphere (Montalb\\'{a}n et al.\\ 2004). Future work must involve not only increased physical realism for the models, but also quantitative comparisons with observations. The methodology outlined in this paper should be applied to a set of real stars, rather than to the idealized evolutionary sequence of representative parameters. Measured stellar masses, radii, and $T_{\\rm eff}$ values, as well as accretion rates, magnetic field strengths, and hot spot filling factors ($\\delta$), should be used as constraints on individual calculations for the stellar wind properties. It should also be possible to use measured three-dimensional magnetic fields (e.g., Donati et al.\\ 2007; Jardine et al.\\ 2008) to more accurately map out the patterns of accretion stream footpoints, wave fluxes, and the flux tubes in which stellar winds are accelerated. This work helps to accomplish the larger goal of understanding the physics responsible for low-mass stellar outflows and the feedback between accretion disks, winds, and stellar magnetic fields. In addition, there are links to more interdisciplinary studies of how stars affect objects in young solar systems. For example, the coronal activity and wind of the young Sun is likely to have created many of the observed abundance and radionuclide patterns in early meteorites (Lee et al.\\ 1998) and possibly affected the Earth's atmospheric chemistry to the point of influencing the development of life (see, e.g., Ribas et al.\\ 2005; G\\\"{u}del 2007; Cuntz et al.\\ 2008). The identification of key physical processes in young stellar winds is important not only for understanding stellar and planetary evolution, but also for being able to model and predict the present-day impacts of solar variability and ``space weather'' (e.g., Feynman \\& Gabriel 2000; Cole 2003)." }, "0808/0808.2585_arXiv.txt": { "abstract": "The evolution of star clusters in the Magellanic Clouds has been the subject of significant recent controversy, particularly regarding the importance and length of the earliest, largely mass-independent disruption phase (referred to as `infant mortality'). Here, we take a fresh approach to the problem, using a large, independent, and homogeneous data set of $UBVR$ imaging observations, from which we obtain the cluster age and mass distributions in both the Large and Small Magelanic Clouds (LMC, SMC) in a self-consistent manner. We conclude that the (optically selected) SMC star cluster population has undergone at most $\\sim$30\\% (1$\\sigma$) infant mortality between the age range from about 3--10 Myr, to that of approximately 40--160 Myr. We rule out a 90\\% cluster mortality rate per decade of age (for the full age range up to $10^9$ yr) at a $>$6$\\sigma$ level. Using a simple approach, we derive a `characteristic' cluster disruption time-scale for the cluster population in the LMC that implies that we are observing the {\\it initial} cluster mass function (CMF). Preliminary results suggest that the LMC cluster population may be affected by $<10$\\% infant mortality. ", "introduction": "One of the most important diagnostics used to infer the formation history, and to follow the evolution of an entire star cluster population is the `cluster mass function' (CMF; i.e., the number of clusters per constant logarithmic cluster mass interval, ${\\rm d}N/{\\rm d}\\log m_{\\rm cl}$). The {\\it initial} cluster mass function (ICMF) is of particular importance. The debate regarding the shape of the ICMF, and of the CMF in general, is presently very much alive, both observationally and theoretically. This is so because it bears on the very essence of the star-forming processes, as well as on the formation, assembly history, and evolution of the clusters' host galaxies on cosmological time-scales. Yet, the observable at hand is the cluster {\\it luminosity} function (CLF; i.e., the number of objects per unit magnitude, ${\\rm d}N/{\\rm d}M_V$). The discovery of star clusters with the high luminosities and the compact sizes expected for (old) globular clusters (GCs) at young ages facilitated by the {\\sl Hubble Space Telescope (HST)} has prompted renewed interest in the evolution of the CLF (and CMF) of massive star clusters. Starting with the seminal work by Elson \\& Fall (1985) on the young Large Magellanic Cloud (LMC) cluster system (with ages $\\lesssim 2 \\times 10^9$ yr), an ever increasing body of evidence seems to imply that the CLFs of young massive clusters (YMCs) are well described by a power law of the form ${\\rm d}N \\propto L^{1+{\\alpha}} {\\rm d}\\log L$, equivalent to a cluster luminosity spectrum ${\\rm d}N \\propto L^{\\alpha} {\\rm d} L$, with a spectral index $-2 \\lesssim \\alpha \\lesssim -1.5$ (e.g., Whitmore \\& Schweizer 1995; Elmegreen \\& Efremov 1997; Miller et al. 1997; Whitmore et al. 1999; Whitmore et al. 2002; Bik et al. 2003; de Grijs et al. 2003; Hunter et al. 2003; Lee \\& Lee 2005; see also Elmegreen 2002). Since the spectral index, $\\alpha$, of the observed CLFs resembles the slope of the high-mass regime of the (lognormal) old GC mass spectrum ($\\alpha \\sim -2$; McLaughlin 1994), this observational evidence has led to the popular theoretical prediction that not only a power law, but {\\it any} initial CLF (and CMF) will be rapidly transformed into a lognormal distribution because of (i) stellar evolutionary fading of the lowest-luminosity (and therefore lowest-mass, for a given age) objects to below the detection limit; and (ii) disruption of the low-mass clusters due to both interactions with the gravitational field of the host galaxy, and internal two-body relaxation effects leading to enhanced cluster evaporation (e.g., Elmegreen \\& Efremov 1997; Gnedin \\& Ostriker 1997; Ostriker \\& Gnedin 1997; Fall \\& Zhang 2001; Prieto \\& Gnedin 2007). However, because of observational selection effects it is often impossible to probe the CLFs of YMC systems to the depth required to fully reveal any useful evolutionary signatures. As such, the young populous cluster systems in the Magellanic Clouds are as yet the best available calibrators for the canonical young CLFs that form the basis of most theoretical attempts to explain the evolution of the CLF and CMF.\\footnote{We point out, however, that with the latest {\\sl HST}/Advanced Camera for Surveys observations of the Antennae interacting galaxies, we are now finally getting to the point where the young (power-law) CLF shape appears to be confirmed independently down to sufficient photometric depths and for larger samples containing more massive clusters than in the Magellanic Clouds (B. Rothberg, priv. comm.).} It is therefore of paramount importance to understand the Magellanic Cloud cluster systems in detail. ", "conclusions": "" }, "0808/0808.3898_arXiv.txt": { "abstract": "Cosmological simulations consistently predict specific properties of dark matter halos, but these have not yet led to a physical understanding that is generally accepted. This is especially true for the central regions of these structures. Recently two major themes have emerged. In one, the dark matter halo is primarily a result of the sequential accretion of primordial structure (ie `Nature'); while in the other, dynamical relaxation (ie `Nurture') dominates at least in the central regions. Some relaxation is however required in either mechanism. In this paper we accept the recently established scale-free sub-structure of halos as an essential part of both mechanisms. Consequently; a simple model for the central relaxation based on a self-similar cascade of tidal interactions, is contrasted with a model based on the accretion of adiabatically self-similar, primordial structure. We conclude that a weak form of this relaxation is present in the simulations, but that is normally described as the radial orbit instability. ", "introduction": "In Henriksen (2006b, H06; 2007, H07) a theory of dark matter relaxation has been proposed that is based on a temporally convergent series solution for the Distribution Function (true phase-space density or DF). The relaxation is effected in a non-mechanistic fashion by maximizing the local Boltzmann function calculated from this DF. This procedure determines all of the parameters in the DF and allows the density distribution $\\rho(r)$ and the pseudo phase-space density $\\phi(r)\\equiv \\rho(r)/\\sigma(r)^3$ to be calculated, as well as any other quantity such as specific angular momentum. Studying these results one finds that, starting from the position and corresponding slopes found in the simulations, either the density flattens rapidly or the pseudo density steepens rapidly within the next decade of smaller radius. There is at present no convincing evidence for either of these trends in the simulations, although unfortunately this region is at or near the current resolution limit. This paper attempts to understand the proposed relaxation more intuitively by suggesting a simple mechanistic model. Such a view may be briefly summarized as `nurture'. In Salvador-Sol\\'e et al. (2007, S07) a rather different explanation for the simulated structures is proposed. In effect one uses the Press-Schechter (1974) formalism as expressed by Lacy and Cole (1993) to calculate the instantaneous merger rate. Since most of the halo mass is added by many small objects, this is approximated by being smooth. This smoothness allows one to accrete the dark matter halo from post-recombination large scale structure in a sort of `layer cake' fashion, where each layer may be deduced from the accretion flow at the epoch when it was added. The innermost layers correspond to the earliest times so that `inside-out' growth is established. The predicted density profile compares well to NFW (Navarro, Frenk and White, 1997; Navarro et al. 2004) profiles for various masses over appropriate ranges and even better over the whole range of scales to an Einasto (Navarro et al., 2004) or S\\'ersic profile (e.g. S07). The latter profile has a finite density at the centre but an infinite slope. Similar comparisons may be made for velocity dispersion and angular momentum (Gonzales-Casado et al., 2007). Such a view may be briefly summarized as `Nature', although there is also a kind of relaxation that leads to adiabatic self-similarity (e.g. Henriksen 2006b, H07) in the asumption of ``smoothness''. Unless it is also hidden in the asumption of `smoothness', the latter picture leaves little room for `thermodynamic relaxation' (ie maximum entropy) as an element in the simulations. Nevertheless, whether or not such relaxation is present in reality in order to explain possible density cores remains an open question. Moreover the remarkable results of the `Via Lactea' simulation (Diemand, Kuhlen and Madau, 2006-DKM06; Madau, Diemand, and Kuhlen, 2008-MDK08) show elaborate sub-structure that is characterized by a definite mass spectrum. Such a hierarchy should be interacting tidally and by dynamical friction and so it provides a mechanism for relaxation (see e.g. the discussions in El-Zant et al., 2004, H07, H06 and Henriksen, 2006a). One indication of the possible presence of such relaxation has been provided by Hoffman et al. (2007). They observe that the halo surface density at the NFW scale radius $r_s$ (that is $\\rho_sr_s$) remains constant during the evolution of an individual halo. This together with virialization within $r_s$ allows them to conclude that $\\phi_s\\propto r_s^{-5/2}$. Some time ago, virialization together with strong dynamical inter-scale coupling were shown to be equivalent to constant surface density and virialization throughout a cascade of structures. This was in a completely different context (Henriksen and Turner, 1984; Henriksen, 1991 and references therein) where it was studied in the context of star formation. It was proposed that a hierarchy of molecular clouds slowly evolved by collisional and tidal interactions into the stellar Initial Mass Function (IMF) as the result of a kind of `ballistic turbulence'. The observation of Hoffman et al (2007) together with the power law substructure (DKM06) suggests a similar argument may apply to dark matter halos. In the stellar case it has been difficult to find any evidence for such an initial fragmented state, and so ironically the picture may be more relevant to dark halos than it is to luminous stars! However Hoffman et al. have not applied the idea of a dynamical cascade to all scales below $r_s$. This will be the subject of the next section. In the third section we give a naive calculation that imitates the layer cake approach of (S07), by simply assuming adiabatic self-similarity as dictated by the primordial perturbation spectrum. Finally in the conclusions we discuss the liklihood of either sort of relaxation. We conclude that a weak form of cascade relaxation is present in the simulations that is not distinguishable from the radial orbit instability. We conclude further that the pure cascade evolution should lead to a sharp break in the density and pseudo-density trends as simulated currently in the next decade or so of resolution. Otherwise we are left only with the weak form of relaxation that is described in this paper as either `nature' or adiabatic self-similarity, and which seems to be equivalent to the radial orbit instability. ", "conclusions": "The preceding sections allow us to conclude that the simulated dark matter halos are, up to the presently available resolution, consistent with the primarily `nature' (S07) adiabatically self-similar smooth growth from the CDM spectrum (e.g. \\ref{fig:nature}). In our semi-analytic treatment of the previous section, we imitate this growth using the notion of adiabatic self-similarity (Henriksen, 2006b) to connect it to the CDM spectrum. In S07 however this was done using the assumption of smooth inside-out or `layer-cake' growth according to the Lacey and Cole (1993) prescription. This employs a Press-Schecter (1974) type argument for the CDM spectrum, together with the statistical treatment of Bond et al. (1991). We conclude that this may be done more simply using the notion of adiabatic self-similarity, which does involve a measure of relaxation. Moreover, although it happens slowly, this approach predicts a central core ($n\\rightarrow -3$) rather than a cusp. What we have termed variously `nurture',`thermodynamic' or `maximum entropy' relaxation based on a cascade of interacting structure (H07 and above) agrees on the qualitative trends with the `nature' view . However it evolves at small scales towards a flat density and a steep pseudo-density more rapidly than is found either in the `nature' discussion or in the simulations if it is assumed to extend close to $r_s$ (\\ref{fig:evolve}). This discrepancy is less pronounced when the relaxation extends only to smaller scales, however (e.g. figure(\\ref{fig:K2evolve})). If we believe that cascade relaxation is playing a r\\^ole in the halo evolution, then equation (\\ref{beta}) and the initial $\\beta\\approx -1$ tell us that the relaxation should indeed be most effective at small scales initially. In time, as $\\beta$ becomes positive, the relaxation should move to large scales in agreement with equation(\\ref{tJ}). The only relaxation that is visible in the simulations at present is in the density profile within two decades or so of $r_s$. The phase space pseudo-density shows no relaxation over this same range. This behaviour can be fitted over this limited range by either of the above relaxation mechanisms, but both mechanisms predict stronger small scale relaxation for which there is as yet no evidence in the simulations. One wonders whether the adiabatic self-similar relaxation that seems to fit the current simulations well, is in fact due to a form of cascade relaxation. Recently a strong case has been made (MacMillan, Widrow and Henriksen, 2006 (MWH06) and references therein) that the relaxation in this region near $r_s$ is due to the Radial Orbit Instability (ROI), so we should reconsider this in the present context. The onset of this instability was found in the MWH06 paper to coincide with the development of the mean square specific angular momentum at $r$ into the Keplerian form $\\propto GM(r)r$. Let us suppose that the angular momentum perpendicular to the radial direction takes the form $\\ell_r^2\\sigma^2(\\ell_r)$, where we use the same notation as in equation (\\ref{coupling}). Then the onset of the ROI is marked by the condition $\\ell_r^2\\sigma^2(\\ell_r)\\approx GM(r)r$ which on taking $\\ell_r\\approx r$ becomes essentially the cascade coupling condition of equation (\\ref{coupling}). The explanation offered in the MWH06 paper is in terms of a resonant interaction between a `bar-like' density perturbation and `particle' (perhaps a sub-structure) orbits. In the present context we may see this as the largest asymmetric sub-structure at $r$. We expect it to be composed of sub-structures and to be interacting with such objects in the environment. It s interesting to note incidentally that this interaction is a kind of coupling between the radial infall and the transverse orbital motion, which delivers free-fall energy to the top of the cascade. The Cascade is really gravitationally driven turbulence. Ome might say that the mode $k\\approx 1/r$ parallel to the infall is coupling to the transverse mode $k_r\\approx 1/\\ell_r$ perpendicular to the radial infall to produce this turbulence. Of course these are all speculative, order of magnitude, arguments. Nevertheless we are inclined to suggest that the simulated halos, hence also the adiabatic self-similarity relaxation that works well for the current simulations, is due to the ROI. This is in turn a mechanism for the conversion of radial infall into transverse orbital motion at $r$. This transverse mode represents the top of the cascade at each radius. We thus identify the ROI with a weak form of cascade relaxation at large scales. This weakness is in accord with equations (\\ref{beta}) and (\\ref{tJ}) that both predict ($\\beta<0$ in the first case) stronger relaxation at small scales. The stronger cascade is expected to be at scales below the resolution of the current simulations. We therefore conclude that cascade evolution may be at work both in reality and in the simulations. We conclude tentatively that `Nature' type adiabatic self-similarity is the weak form of the cascade relaxation and that it is equivalent to the ROI. Our predictions that would confirm the presence of cascade evolution are that there is a halo density core rather than a cusp and that the pseudo-density power law should break at higher resolution in both scale and mass. In the Via Lactea run this region is just inside their reported convergence radius at about 1kpc. There may be weak evidence in figure 1 of DKM06 that a density flattening is not excluded. Should this density and pseudo-density behaviour not appear down to say 100pc in a Milky Way type halo, then the strong cascade relaxation is not present in the simulations. It is then probably not present in reality, unless for some reason the necessary interactions at a distance between sub-structures is discriminated against in the simulations. DKM06 do observe that at lower resolution than that of the Via Lactea run, the small sub-halos are poorly resolved." }, "0808/0808.3917_arXiv.txt": { "abstract": "{It has been reported that exoplanet host stars are lithium depleted compared to solar-type stars without detected massive planets. } {We investigate whether enhanced lithium depletion in exoplanet host stars may result from their rotational history. } {We have developed rotational evolution models for slow and fast solar-type rotators from the pre-main sequence (PMS) to the age of the Sun and compare them to the distribution of rotational periods observed for solar-type stars between 1~Myr and 5~Gyr.} {We show that slow rotators develop a high degree of differential rotation between the radiative core and the convective envelope, while fast rotators evolve with little core-envelope decoupling. We suggest that strong differential rotation at the base of the convective envelope is responsible for enhanced lithium depletion in slow rotators. } {We conclude that lithium-depleted exoplanet host stars were slow rotators on the zero-age main sequence (ZAMS) and argue that slow rotation results from a long lasting star-disk interaction during the PMS. Altogether, this suggests that long-lived disks ($\\geq$ 5~Myr) may be a necessary condition for massive planet formation/migration.} ", "introduction": "As the discovery rate of exoplanets has steadily increased over the years, now reaching nearly 300 detections, interest has grown in the properties of their host stars (Santos et al. 2003; Kashyap et al. 2008). Israelian et al. (2004) report that solar-type stars with massive planets are more lithium-depleted than their siblings without detected massive planets. This result has recently been confirmed by Gonzalez (2008). Lithium over-depletion in massive exoplanet hosts appears to be a generic feature over a restricted T$_{eff}$ range from 5800 to 5950~K, independent of the planet's orbital properties. Metallicity effects and/or the early accretion of planetesimals have been ruled out as the origin of lithium over-depletion in exoplanet hosts (Castro et al. 2008). Instead, it has been suggested that enhanced lithium burning could stem from the tidal effect of the giant planet on the host star, as it induces rotationally-driven mixing in the stellar interior (Israelian et al. 2004). However, the mass of giant exoplanets is much lower than the mass of the convective envelope of solar-type stars ($\\geq$0.02$M_\\odot$). In addition, enhanced lithium depletion is seen for stars with giant planets that span a wide range of semi-major axes, up to about 2~AU. Whether the tidal interaction between a distant massive planet and the host star can be strong enough to induce rotationally-driven mixing remains to be investigated. Rotationally-driven mixing may instead be related to the rotational history of the star (Takeda et al. 2007; Gonzalez 2008). We revisit here the link between lithium depletion and the rotational history of solar-type stars, and discuss its relationship with massive planet formation. In Section 2, we present models for the rotational evolution of solar-type stars from the early PMS to the age of the Sun. We use these models to fit the most recent measurements of rotational periods obtained for solar-type stars in star forming regions and young open clusters. We discuss the properties of 2 extreme models: one reproduces the evolution of fast rotators over time, and the other the evolution of slow rotators. We find that slow rotators develop a high degree of differential rotation between the inner radiative zone and the outer convective envelope, while fast rotators evolve with little core-envelope decoupling. From these results, we investigate in Section 3 the evolution of lithium abundances in slow and fast rotators. We suggest that the large velocity shear at the core/envelope interface in slow rotators leads to more efficient Li-depletion. In Section 4, we relate the Li-depletion pattern of massive exoplanet host stars to their rotational history. We argue that both are dictated by their initial angular momentum and the lifetime of their protoplanetary disk. We conclude that long-lived disk may be a necessary condition for massive planet formation/migration around young solar-type stars. ", "conclusions": "Based on what we currently know of the rotational properties of young stars, of the lithium depletion process in stellar interiors, and of the angular momentum evolution of solar-type stars, it seems likely that the lithium-depleted content of massive exoplanet host stars is a sequel to their specific rotational history. This history is predominantly dictated by star-disk interaction during the PMS. Using simple rotational models, it is suggested here that the long disk lifetimes of initially slow rotators lead to the development of a large velocity shear at the base of the convective zone. The velocity gradient in turn triggers hydrodynamical instabilities responsible for enhanced lithium burning on PMS and MS evolution timescales. Rotationally-driven lithium depletion in exoplanet host stars can thus be at least qualitatively accounted for by assuming PMS disk lifetimes of 5-10 Myr. Such long-lived disks may be a necessary condition for planet formation and/or migration around young solar-type stars, at least for the class of giant exoplanets detected so far (Ida \\& Lin 2008). Admittedly, the scenario outlined here is based on rotational models that do not incorporate a detailed physical description of the processes at work. The conclusion therefore remains qualitative and has to be ascertained by the development of more sophisticated angular momentum evolution models. Most important, the transport processes that lead to a strong core-envelope coupling in fast rotators and a weak one in slow rotators still have to be identified. While this goal is well beyond the scope of the present study, we hope that the simplified models presented here may provide useful guidelines for further refinements." }, "0808/0808.3310_arXiv.txt": { "abstract": "The asymmetric time dependence and various statistical properties of polarity reversals of the Earth's magnetic field are utilized to infer some of the most essential parameters of the geodynamo, among them the effective (turbulent) magnetic diffusivity, the degree of supercriticality, and the relative strength of the periodic forcing which is believed to result from the Milankovic cycle of the Earth's orbit eccentricity. A time-stepped spherically symmetric $\\alpha^2$-dynamo model is used as the kernel of an inverse problem solver in form of a downhill simplex method which converges to solutions that yield a stunning correspondence with paleomagnetic data. ", "introduction": "The hydromagnetic dynamo in the Earth's outer core converts gravitational and thermal energy into magnetic energy \\cite{MERRILL}. One of the most impressive feature of the geomagnetic field is the irregular occurrence of polarity reversals. Averaged over the last few million years the mean rate of reversals is approximately 4-5 per Myr, although the last reversal occurred approximately 780000 years ago. At least two, but very likely three \\cite{GALLET,COURTILLOT} superchrons have been identified as ''quiet'' periods of some tens of millions of years showing no reversal at all. Knowledge on ancient magnetic field data is mainly obtained from paleomagnetic measurements of permanent magnetization from (frozen) lava and sedimentary rocks. However, appropriate paleomagnetic sites are rare, unevenly distributed across the Earth's surface and, furthermore, actual dating methods allow only a rather rough time-resolution. Hence only few reversal characteristics can be evaluated as robust \\cite{MERRILL2}. One of the commonly accepted features of reversals is their pronounced temporal asymmetry with the initial decay of the dipole being much slower than the subsequent recreation of the dipole with opposite polarity \\cite{VALET}. Recent numerical simulations have been successful in reproducing not only the dominance of the axial dipole and the spectrum of the geomagnetic field, but also the irregular occurrence of polarity reversals \\cite{ROGLA,WICHTOLSON,AUBERT}. Polarity reversals were also observed in one \\cite{BERHANU} of the recent liquid sodium dynamo experiments which have flourished during the last decade \\cite{RMP}. It is important to note, however, that neither in simulations nor in experiments it is possible to accommodate all dimensionless parameters of the geodynamo \\cite{GUBBINS,ZAMM}, and many of them are not even well known \\cite{ROGLA2}. A complementary way to acquire knowledge about the geodynamo is to use available magnetic field data for constraining the source of dynamo action. Most famous among those attempts is the frozen-flux approximation \\cite{ROSCO,HOLMEOLSEN} which allows to infer the tangential flow at the core-mantle-boundary from secular variation of the geomagnetic field. Going beyond this frozen-flux approximation and trying to infer properties of the geodynamo in the depth of the Earth's core has been less successful so far. Such trials to ''look inside the dynamo'' by utilizing spectral data worked nicely for simplified dynamo models \\cite{AN,PEPI,PRE} but are hardly applicable for real world dynamos. With this sobering experience in mind, in the present paper we undertake a rather uncommon attempt to constrain some of the most significant parameters of the geodynamo by various characteristics of paleomagnetic reversal records. Most prominent among those characteristics is the above mentioned temporal asymmetry of reversals. Two further characteristics reflect some sort of ordering in the otherwise irregular reversal sequence. The first one is the clustering property of reversals which was discovered only recently \\cite{CARBONE,LUCA}. Clustering of reversals is manifested in an enhanced probability of a consecutive reversal shortly after a first reversal has occurred, which results in a deviation from the Poisson distribution that would hold for uncorrelated events. The second one is the appearance of a $\\sim$100 kyr periodicity in the distribution of the residence times (RTD) between reversals which is believed to result from the Milankovic cycle of the Earth's orbit eccentricity \\cite{CONSOLINI,LORITO}. Based on these three input features we will examine in this paper a simplifying $\\alpha^2$-model of the geodynamo for which we estimate the degree of supercriticality, the noise level, the relative strength of the periodic forcing, and the effective (turbulent) magnetic diffusivity of the Earth's outer core. Actually, a few dependencies on individual parameters were already published in preceeding papers. In \\cite{EPSL} the dependence of typical time scales on the supercriticality of the dynamo was studied, in \\cite{LUCA} some dependencies of the clustering property on the supercriticality and the noise level were shown, and in \\cite{EPJB} the influence of the diffusion time scale on the RTD between reversals was touched upon. What is new in the present paper is that we take all three reversal features together and try to infer from them some essential parameters of the geodynamo. We will start with a presentation of the forward dynamo problem for which we will use a rather simple, spherically symmetric mean-field dynamo of the $\\alpha^2$ type. This simple model had turned out to be quite helpful for understanding the basic principle of the reversal process as a noise-induced relaxation-oscillation in the vicinity of an exceptional point of the spectrum of the non-selfadjoint dynamo operator \\cite{EPSL,PRL,GAFD}. This exceptional point, at which two real eigenvalues coalesce and continue as a complex conjugated pair of eigenvalues, is associated with a nearby local maximum of the growth rate situated at a slightly lower magnetic Reynolds number. It is the negative slope of the growth rate curve between this local maximum and the exceptional point that makes stationary dynamos vulnerable to noise. Then, the instantaneous eigenvalue is driven towards the exceptional point and beyond into the oscillatory branch where the sign change of the dipole polarity happens. After having delineated the simplified mean-field dynamo model we will present the solution method for the inverse problem and the main results. The paper concludes with a summary and a speculation on the possible consequences of our findings for the general understanding and the numerical simulations of the geodynamo. ", "conclusions": "With version 2, we have obtained the best reproduction of the paleomagnetic input data for a 10 times supercritical dynamo, a relative strength of the periodic forcing of some 15 per cent, and an effective magnetic diffusion time of approximately 65 kyr which is by a factor 3.5 smaller than the value that would result from the molecular conductivity. The latter is perhaps the most important result of the inversion, for the following reason: The conductivity enhancing effect of turbulence is a subject of ongoing debate. Only very few estimations exists for the turbulent diffusivities within the Earth's outer core. However, their values are of extraordinary importance for the key parameters of the geodynamo (e.g. Ekman number or the magnetic Prandtl number) as well as for the estimation of turbulent transport properties and the destructive influence of turbulence on magnetic field generation through turbulent field diffusion. The molecular values of the dimensionless parameters are extremely small and cannot be realized in simulations so that usually enhanced values are applied that shall resemble the effective (i.e. turbulent) quantities. A more detailed information on the turbulent diffusivities therefore delivers important information on the significance of the parameter space accessible in global MHD simulations for the geodynamo. Furthermore, reality is more complicated, as it is very likely that a conductivity reduction due to turbulence would be anisotropic, because the small scale turbulence in a fast rotating object like the Earth is subject to preferred directions parallel to the rotation axis (and also along the dominating field component) so that the convection cells (that determine the turbulent transport of physical quantities) are oriented parallel to the rotation axis and elongated along the magnetic field. It is well known that an anisotropic conductivity could have a tremendous effect on the selection between equatorial and axial dipole solution \\cite{TILGNER,ELSTNER}. Roughly speaking, axially aligned rotating columns tend to decrease the conductivity for horizontal currents less than the conductivity for vertical currents, and this effect leads to a preference of the axial dipole compared to the equatorial dipole. This could be an extremely important point since the conclusions of many geodynamo related papers rely on an isotropic conductivity when looking for criteria for the selection of axial or equatorial dipoles. The stunning agreement of paleomagnetic and numerical reversal characteristics as shown in figures 1, 2 and 3 gives support to our hypothesis that reversals are indeed noise triggered relaxation oscillations in the vicinity of an exceptional point of the spectrum of the dynamo operator. In this respect it is important to note that in particular the time asymmetry and the clustering property are intrinsic and robust properties of the model that appear for very wide regions of parameter (if not for all). By solving the inverse problem we have not ''produced'' them, but have only fine-tuned the dynamo parameters to fit optimally the paleomagnetic data. We have carefully tried not to over-interpret our simple model by focusing only on those parameters to be determined, and those functionals to be minimized, that refer to the temporal properties of reversal sequences, and not to any spatial features. This makes us optimistic that the results will prove robust when inversions of this kind will later be repeated using more realistic dynamo models. Given that one downhill simplex run for our simple model takes already one week, one can imagine that corresponding runs with better dynamo models will lead to significant computational costs. \\ack This work was supported by Deutsche Forschungsgemeinschaft in frame of SFB 609." }, "0808/0808.1301_arXiv.txt": { "abstract": "Motivated by the possibility that the fundamental ``constants'' of nature could vary with time, this paper considers the long term evolution of white dwarf stars under the combined action of proton decay and variations in the gravitational constant. White dwarfs are thus used as a theoretical laboratory to study the effects of possible time variations, especially their implications for the future history of the universe. More specifically, we consider the gravitational constant $G$ to vary according to the parametric relation $G = G_0 (1 + t/\\tG)^{-p}$, where the time scale $\\tG$ is the same order as the proton lifetime $t_{P}$. We then study the long term fate and evolution of white dwarf stars. This treatment begins when proton decay dominates the stellar luminosity, and ends when the star becomes optically thin to its internal radiation. ", "introduction": "One of the overarching but unresolved questions in physics is whether or not the fundamental constants are truly constant, or if they could vary with time. A related question is whether these constants could, in principle, have other values in far-away regions of space-time (i.e., in effectively other universes). Current experiments place limits on the time variations of the fundamental constants, e.g., the gravitational constant $G$ and the fine-structure constant $\\aem$ (Uzan 2003 and references therein). These constraints can be expressed in terms of corresponding time scales, $\\tg = G / {\\dot G}$ and $\\tem = \\aem / {\\dot \\aem}$, which are found in the range $10^{10} - 10^{12}$~yr (Uzan 2003). Although these time scales are safely longer than the current age of the universe ($t_0$ = 13.7 Gyr; e.g., Spergel et al. 2007), these values are much shorter than some time scales that are experimentally accessible; as one example, the proton lifetime has a measured lower bound of $10^{33}$ yr (Super-K 1999). As a result, over the vast expanses of time available to the universe in the future, time variations in the physical constants could change the projected future history of the universe. This paper addresses this issue by considering possible time variations of the gravitational constant and its corresponding effects on the long term evolution of white dwarf stars. In this context, we are using white dwarfs as a theoretical laboratory to consider the effects of time-varying $G$. This choice is justified because white dwarfs are among the simplest and hence most well understood stellar objects (starting with Chandrasekhar 1939) and because they play an important role in the future history of the universe (Adams \\& Laughlin 1997, hereafter AL97; see also Cikovic 2003, Dyson 1979, Islam 1977). For example, almost all stars turn into white dwarfs after their nuclear burning phase, and hence a sizable fraction of the accessible baryonic content of the universe is locked up in white dwarfs after stellar evolution has run its course (for completeness, note that a sizable fraction of the baryons remain in the medium between galaxies in dusters, e.g. Nagamine \\& Loeb 2004). To address this issue, we must define both the time scales and the functional form of the time variations under consideration. Although an enormous variety of time variations are possible, we narrow the scope by allowing the gravitational constant $G$ to vary according to the parametric relation \\be G(t) = G_0 (1 + t/\\tG)^{-p} \\, , \\ee where $G_0$ is the current value, and where $\\tG$ and $p$ are parameters. This functional form is perhaps the simplest type of allowed time variation --- the value of $G$ reduces to the current value at the current cosmological epoch and decreases as a power-law in the long term future. Note that current experimental limits imply that $\\tg = G/{\\dot G} > 10^{12}$ yr (e.g., Table IV of Uzan 2003; see also Barrow 1996) so that $\\tG \\gg t_0$. In the long term future, proton decay is one of the most important energy sources for white dwarfs. This problem thus has two time scales of interest: the proton lifetime $t_P$ and the time scale $\\tG$ for variations in $G$. In the limit $t_\\star \\gg t_P$, white dwarf evolution closely follows previous results obtained using a fixed gravitational constant (AL97, Adams et al. 1998; Dicus et al. 1982). As a result, this paper primarily considers the limit $t_\\star \\ll t_P$, where white dwarf evolution is dominated by changes in gravity, and the case where $\\tG$ is comparable to $t_P$. Although this work will focus on time variations in the gravitational constant $G$, for completeness we note that time variations in other physical quantities are possible. For example, the masses of the fundamental particles or, equivalently, the ratio of electron to proton mass $\\mu = m_e / m_P$, could be time dependent (Calmet \\& Fritzsch 2006). We note that the quark masses, which play a role in determining the proton mass, are actually the fundamental constants of the underlying theory; however, the ratio $\\mu$ has been experimentally constrained (see Uzan 2003 and references therein). In this paper we consider the variation of $G$ by itself in order to isolate the effects of its possible time evolution. However, fundamental theories (string theory, M theory) suggest that the forces are unified and hence may vary in strength in a coupled manner (Uzan 2003 provides a brief review of the possibilities). Unfortunately, we do not have a definitive working theory of how such coupled time variations are expected to occur, and we leave such a more comprehensive treatment for future work. This paper is organized as follows. In \\S 2, we outline a basic model for white dwarf structure, including our treatment of time variations in $G$ and proton decay. One interesting complication is that proton decay leads to a downward cascade for the atomic weights of the constituent nuclei; this issue is addressed in some detail. In \\S 3 we outline the basic results of the paper, including the H-R diagrams for the long term evolution of these degenerate stars and their corresponding chemical evolution. We then conclude in \\S 4 with a summary and discussion of our results. ", "conclusions": "This paper has constructed a model for the structure and evolution of white dwarf stars under the action of both proton decay and time variations in the gravitational constant (section 2). This model includes variations in chemical composition (Figure 2 and section 2.3) which affects the equation of state. The model also includes separate accounting for the inner degenerate regions (which obey the polytropic equation of state [1]) and the outer non-degenerate regions. Our main findings can be summarized as follows: In the absence of baryon decay, white dwarfs grow larger as the strength of gravity decreases (Figure 4). The radius reaches a maximum value when the entire star becomes non-degenerate; at this time, the radius is $\\sim 100$ times larger than its starting value. With the inclusion of baryon decay, white dwarfs lose mass with time and follow tracks in the H-R diagram as shown in Figure \\ref{fig:HRC}. In all cases, after an initial transient phase, the stars grow dimmer and redder with time, and hence move to the lower right in the H-R diagram. While the stars remain (primarily) degenerate, the tracks have a shallow slope, with $L \\propto T^{12/5}$. After the stars lose enough mass to become primarily non-degenerate, the tracks become much steeper, with $L \\propto T^{12}$. These slopes would be exact in the absence of changes in the chemical composition; the chemical variations and the time variations in $G$ lead to small departures from this behavior. During the degenerate regime of evolution, the radii of these white dwarf stars grow larger due to mass loss from proton decay and due to the weakening of gravity. During the later, non-degenerate phase, the radii become smaller. The maximum value of the radius occurs near the transition between the degenerate and non-degenerate regimes, with a value that is typically $\\sim 10$ times the starting radius (see Figure \\ref{fig:RvstC}). In addition to understanding the long term fate and evolution of white dwarfs, one motivation for this work is to understand the larger issue of the future of the universe. Previous projections of the future (e.g. AL97, Dyson 1979) are predicated on the assumption that the laws of physics are known and will not change with time. One important issue is thus the question of whether or not the constants of nature change with time, and how such variations would change projections of our future history. In the case of white dwarf evolution, we find that time variation leads to quantitative --- but not qualitative --- changes in the future timeline. Time varying $G$ leads to variations in the exact mass, size, and composition of white dwarfs as they become non-degenerate rock-like objects. However, the starting states and final states are essentially the same, so time variations in gravity do not lead to fundamental changes in the overall picture. White dwarfs start as degenerate objects with the current value of $G$, and hence are constrained to begin their evolution in nearly the same states. At the other end of time, white dwarfs cease to act as stars when they become optically thin to their internal radiation. At this epoch, the ``stars'' are essentially large rocks of Hydrogen ice, and their structure is non-degenerate and largely independent of gravity. As a result, time variations in $G$ primarily affect the intermediate states, and this work shows that the modifications are relatively modest (see Figures \\ref{fig:rhovtC} - \\ref{fig:HRC})." }, "0808/0808.1889_arXiv.txt": { "abstract": "Some early-type galaxies show a correlation between color and integrated magnitude among the brighter metal-poor globular clusters (GCs). This phenomenon, known as the blue tilt, implies a mass-metallicity relationship among these clusters. In this paper we show that self-enrichment in GCs can explain several aspects of the blue tilt, and discuss predictions of this scenario. ", "introduction": "Several recent photometric studies of extragalactic globular cluster (GC) systems have discovered a novel feature. The blue, metal-poor population of GCs shows a color-magnitude relation at bright magnitudes: the more luminous GCs are redder. This ``blue tilt\" was first noticed in the GC systems of massive elliptical (E) galaxies (Harris \\etal~2006; Strader \\etal~2006), but has since been seen in the Sa/S0 NGC 4594 (Spitler \\etal~2006), fainter early-type galaxies in Virgo (Mieske \\etal~2006), and even in normal spirals (Strader \\etal~2008, in preparation). Figure 1 shows the blue tilt in M87 (Strader \\etal~2006; see also Mieske \\etal~2006). The blue tilt is common but not ubiquitous---the massive E NGC 4472 does not have it, for example, and among galaxies that do show a tilt, the slope varies. There is no evidence for a tilt in the red, metal-rich subpopulation of GCs. Many explanations for this phenomenon have been proposed. These include (i) self-enrichment, (ii) ``pollution\" of the GC sequence with objects such as dwarf galaxy nuclei or merged star clusters (e.g., Bekki \\etal~2007), and (iii) gravitational capture of metal-rich field stars by GCs (Mieske \\etal~2006). A number of other scenarios are discussed in Mieske \\etal~(2006) but are thought by them to be much less probable. Option (ii) is considered by both Mieske \\etal~(2006) and Strader \\etal~(2006), who conclude that it is unlikely. While the nuclei of Virgo dwarfs show their own color-magnitude relation (Lotz \\etal~2001), the red mean colors of the luminous GCs are not due solely to the addition of bright red objects, but instead come from systematic shifts in the GC color distribution with magnitude. In addition, the number of nuclei needed is much larger than the handful of stripped nuclei that have been found in either the Virgo or Fornax clusters (see also discussion in \\S 4). Except for a few objects, the sizes of the blue tilt GCs are generally small, inconsistent with the expectations of merged clusters. Merged clusters should be larger as the orbital energy of the binary cluster is converted into kinetic energy in the merger product. The best candidate for a merged cluster, NGC 1846 in the Large Magellanic Cloud (Mackey \\& Broby Nielsen 2007), has an unusually large projected half-light radius of $\\sim 15$ pc (D.~Mackey, private communication). While it seems likely that such exotic objects contribute to the GC systems of massive galaxies at some level, especially at the brightest magnitudes (Harris \\etal~2006; Strader \\etal~2006), they are unlikely to be primarily responsible for the blue tilt. Option (iii) was investigated using collisional N-body simulations by Mieske \\& Baumgardt (2007). They report that field star capture is inefficient even in optimistic circumstances, and conclude that this is not the cause of the blue tilt. Our working assumption is that self-enrichment remains the most likely explanation for the blue tilt. In this case the basic explanation is that the color-magnitude relation reflects a physical mass-metallicity relation for metal-poor GCs. As discussed in Strader \\etal~(2006), there are a variety of GC self-enrichment models and simulations in the literature (e.g., Morgan \\& Lake 1989; Brown \\etal~1991; Parmentier \\etal~1999; Parmentier \\& Gilmore 2001; Parmentier 2004; Recchi \\& Danziger 2005). Here we refer to models that focus on self-enrichment as the primary determinant of cluster metallicity, and not to the larger body of work on the origin of abundance anomalies in Galactic GCs that primarily involve light elements such as C, N, O, and Mg. Among GC self-enrichment models, there is no consensus on the dominant physical processes or expected level of enrichment as a function of cluster mass. The lack of a tilt in the red GCs might be expected if the self-enriched metals were a constant fraction of the stellar mass, since the overall metal content of red GCs is a factor of $\\sim 10$ higher than the blue GCs (Strader \\etal~2006; Mieske \\etal~2006). We illustrate this point further in \\S 3. A possible relationship between GC self-enrichment and a mass-metallicity relation among GCs was discussed for the oldest halo Galactic GCs by Parmentier \\& Gilmore (2001), prior to the discovery of the blue tilt in external galaxies. However, Parmentier \\& Gilmore were attempting to explain an anticorrelation between GC metallicity and mass, the opposite of the blue tilt. Here we do not attempt detailed modeling of the self-enrichment process (see the studies cited in the previous paragraphs for such models). Our model differs from many of those previously cited in that we invoke a single generation of star formation; within the context of self-enrichment, we expect a spread in metallicity in individual star clusters. This is inconsistent with the chemical homogeneity observed in most Galactic GCs, but the Galaxy does not have a blue tilt (see additional discussion in \\S 8). Our approach is to use scaling relation assumptions about the putative enrichment to explore the conditions under which the observed GC mass-metallicity relationships may be reproduced. These conditions may then be used as additional constraints on GC formation. ", "conclusions": "We have examined a number of scenarios for the blue tilt, a mass-metallicity relationship for metal-poor GCs observed in many external galaxies. The model that seems to best reproduce current observations is self-enrichment in proto-GC clouds; in this model, star formation is governed by supernova feedback, and the efficiency is a function of the protocloud mass. Certain efficiency scalings produces no mass-radius relation for the resultant GC (in accordance with observations), and are equivalent to an energy balance condition for star formation. We have also speculated as to how the observed variations in the slope of the blue tilt and the break mass might be reproduced. Contrary to the comments in Strader \\etal~(2006), a dark matter halo is not necessarily required for effective self-enrichment, at least given the assumptions of the current model. Many of the color-metallicity relations used are uncertain, and may be the source of some of the observed scatter in blue tilt slopes. Since it is unlikely that these relations will see significant improvement soon, it would be useful to make photometric observations of different galaxies with a common color (e.g., $B-I$ or $g-z$) so that the intercomparisons are at least consistent. Another route is to directly estimate the metallicities of individual clusters using spectroscopy, or with near-IR imaging, which is more sensitive to metallicity than optical colors. Most of the break masses are poorly determined, so deeper imaging of selected galaxies would be beneficial. The assumption of a variable formation efficiency can be tested directly by measuring cloud and cluster masses in nearby systems of young massive clusters, as suggested by Ashman \\& Zepf (2001). With ALMA, accurate cloud masses and radii can be determined in such systems, testing both the efficiency of cluster formation and the initial mass-radius relationship for clouds as a function of external parameters such as pressure. Self-enrichment has been separately proposed to explain a number of abundance anomalies in Galactic GCs, primarily involving light elements such as C, N, O, and Mg. The culprits are generally argued to be an unknown combination of massive main-sequence stars and intermediate-mass AGB stars. A natural expectation of self-enrichment models for the blue tilt would be the appearance of abundance anomalies associated with blue tilt GCs. The integrated abundances of C, N, and Mg can all be estimated in extragalactic GCs from low-resolution spectra. For example, Cenarro \\etal~(2007) find CN line strengths for blue tilt GCs in NGC 1407 that are much larger than in Galactic GCs at comparable metallicities; unfortunately, less luminous GCs in NGC 1407 have not been observed. Studying the abundance patterns of GCs both above and below the blue tilt break mass could constrain the timing and duration of the hypothesized self-enrichment events. The self-enrichment model makes another prediction that was described in \\S 1 and \\S 3. Unless star formation in the GC is oddly segregated by mass, such that the high mass stars form, explode as supernovae, and enrich the intracluster gas before any low-mass stars form, then the blue tilt GCs would be expected to have internal dispersions in not just the lighter elements such as C through Mg, but also in heavier elements produced by massive star supernovae. In the Milky Way the only GC known to have a large star-to-star spread in the $\\alpha$-elements heavier than Mg is $\\omega$ Cen (e.g., Freeman \\& Rodgers 1975; Smith et al. 2000), a cluster that has been viewed as a test case of self-enrichment and chemical evolution (e.g., Carraro \\& Lia 2000; Ikuta \\& Arimoto 2000; Platais et al. 2003; Marcolini et al. 2007), and possibly M22 (Norris \\& Freeman 1983; Lehnert et al. 1991; Anthony-Twarog et al. 1995). The majority of Galactic GCs are homogeneous to a high degree in the $\\alpha$-elements. However, the Milky Way GC system is thought not to exhibit a blue tilt (although see Parmentier \\& Gilmore 2001), so there is no conflict with the self-enrichment scenario for this phenomenon. However, the challenge in other galaxies could be two-fold: (i) if blue tilt GCs are chemically homogeneous, then the self-enrichment picture may be less compelling; (ii) if blue tilt GCs are typically inhomogeneous in heavy elements, why is this not true in Galactic GCs of comparable mass? Regarding (i), it is intriguing that four of the most massive GCs in M31 have large internal metallicity spreads (Fuentes-Carrera \\etal~2008). Forthcoming homogeneous photometric and spectroscopic datasets should allow the detection of a blue tilt in M31, if it exists. From a theoretical standpoint, self-enrichment models such as those of Brown \\etal~(1991) attempt to reconcile supernova-induced enrichment with chemical homogeneity. With regard to (ii), we suggest reluctantly that it might be necessary to invoke fundamental differences between the parent clouds of Galactic GCs and those in blue tilt galaxies." }, "0808/0808.1071_arXiv.txt": { "abstract": "{The equilibrium rotation of tidally evolved ``Earth-like'' extra-solar planets is often assumed to be synchronous with their orbital mean motion. The same assumption persisted for Mercury and Venus until radar observations revealed their true spin rates. As many of these planets follow eccentric orbits and are believed to host dense atmospheres, we expect the equilibrium rotation to differ from the synchronous motion. Here we provide a general description of the allowed final equilibrium rotation states of these planets, and apply this to already discovered cases in which the mass is lower than 12 $M_{\\oplus}$. At low obliquity and moderate eccentricity, it is shown that there are at most four distinct equilibrium possibilities, one of which can be retrograde. Because most presently known ``Earth-like'' planets present eccentric orbits, their equilibrium rotation is unlikely to be synchronous.} ", "introduction": "After a significant number of discoveries of new extra-solar gaseous giant planets, a new barrier has been passed with the detections of several planets in the Neptune and even Earth-mass ($M_{\\oplus}$) regime: $5-12$ $M_{\\oplus}$ \\citep{Rivera_etal_2005,Lovis_etal_2006,Udry_etal_2007,Bonfils_etal_2007}. If the commonly accepted core-accretion model can account for the formation of these planets, resulting in a mainly icy/rocky composition, the fraction of the residual He-H$_2$ atmospheric envelope accreted during the planet migration is not tightly constrained for planets more massive than the Earth \\citep[e.g.][]{Alibert_etal_2006}. A minimum mass of below 10 $M_{\\oplus}$ is usually considered to be the boundary between terrestrial and giant planets, but \\citet{Rafikov_2006} found that planets more massive than 6 $M_{\\oplus}$ could have retained more than 1 $M_{\\oplus}$ of the He-H$_2$ gaseous envelope. For comparison, masses of Earth's and Venus' atmosphere are respectively $\\sim 10^{-6}$ and $10^{-4}$ times the planet's mass. Despite significant uncertainties, these discoveries of ``super-Earths'' provide an opportunity to test some properties that could be similar to those of our more familiar telluric planets. Because some of these planets are potentially in the ``habitable zone'' \\citep{Udry_etal_2007,Selsis_etal_2007}, their spin state is an important factor in understanding their climate. For planets with a negligible atmosphere such as Mercury, solid tides induced by the host star are expected to despin the planet until an equilibrium position or a capture in a spin-orbit resonance, depending on the orbital eccentricity and the permanent quadrupole moment of inertia \\citep{Goldreich_Peale_1966,Correia_Laskar_2004}. % However, for planets with dense atmosphere such as Venus, thermal atmospheric tides, driven by solar insolation, have a profound influence on the spin, and may destabilize the previous tidal equilibrium state. \\citet{Correia_Laskar_2001,Correia_Laskar_2003I} investigated the combined effect of gravitational and atmospheric tides on Venus and showed that the planet could evolve into two different final states, the existing retrograde motion being the most probable one. This study was performed with the small eccentricity approximation that may no longer be adequate for ``Earth-like'' planets, which exhibit a wide range of eccentricities, orbital distances, or central star types. Following the work of \\citet{Laskar_Correia_2004}, we generalize here these previous studies and investigate the possible equilibrium rotation states of ``Earth-like'' planets in the presence of significant eccentricity. Although our knowledge of these planets is restricted to their orbital parameters and minimum masses, we attempt to place new constraints on their surface rotation rate, assuming that they have a dense atmosphere. ", "conclusions": "We have derived a simple model that permits the determination of the final equilibrium rotation of ``Earth-like'' planets. Our model contains some uncertain parameters related to the dissipation within the planets, but we were able to gather all this information in a single parameter, $\\omega_s$. By varying this, we can then cover all possibilities for the rotation of terrestrial planets. We demonstrated that for a planet of moderate eccentricity and low obliquity, at most four final equilibrium positions are possible. For eccentricities higher than $ e \\sim 0.5 $, terms of higher degree in $e$ should be considered in Eq.(\\ref{eq02}) that may generate additional equilibrium positions. Based on the present rotation of Venus, we provided an estimate of $\\omega_s$ for different environments (Eq.\\,\\ref{eq60}). An important consequence is that the ratio $ \\omega_s / n $ increases rapidly with the semi-major axis and mass of the star because $ \\omega_s /n \\propto (a M_*)^{2.5} $. The effect of the atmosphere on the equilibrium rotation is therefore more relevant for planets that orbit Sun-like stars at not close distances. Unfortunately, this situation is precisely the opposite of that for the discovered Earth-like planets (Table\\,\\ref{Tab1}), since the radial velocity technique is more sensitive to the detection of short-period planets. {\\bfi As a consequence, unlike Venus, none of these planets can be stabilized with a rotation rate $\\omega < n $, because for $ e > 0.1 $ these final states only exist if $ a M_* > 0.2 $\\,AU\\,$M_\\odot$ (Fig.\\ref{Fig2}).} The effect of atmospheric tides is extremely small ($ \\omega_s / n \\sim 0 $) and the equilibrium rotation is essentially driven by the tidal gravitational torque. However, their significant eccentricity ($0.1 < e < 0.2$) moves the equilibrium position away from the synchronous motion. Capture in this resonance is unlikely, but capture in a higher order spin-orbit resonance cannot be rejected, in particular for {\\small GJ}\\,581\\,d and {\\small GJ}\\,674\\,b, depending on the asymmetry of their mass distribution." }, "0808/0808.0077_arXiv.txt": { "abstract": "Strong size and internal density evolution of early-type galaxies between $z\\sim 2$ and the present has been reported by several authors. Here we analyze samples of nearby and distant ($z\\sim 1$) galaxies with dynamically measured masses in order to confirm the previous, model-dependent results and constrain the uncertainties that may play a role. Velocity dispersion ($\\sigma$) measurements are taken from the literature for 50 morphologically selected $0.81.5$ results. Alternatively, the different studies sample galaxies with a wide range in masses, and therefore mass-dependent size evolution could lead to apparent discrepancies among the samples. This is explored in the following section. Our robust results strengthen the results from previous studies. We conclude that early-type galaxies at $z=1$ are $\\sim 2$ times smaller than local early types with the same mass, and that at $z=2-2.5$ this size difference is likely increased to a factor of $\\sim$4, as previously observed by \\citet{zirm07}, \\citet{toft07}, \\citet{vandokkum08b}, and \\citet{buitrago08}. \\subsection{Comparison with Model Predictions}\\label{dis:2} The fact that we see considerable evolution in galaxy size with redshift is not surprising from a theoretical perspective. Most semianalytic models of galaxy formation in a $\\Lambda$CDM universe predict substantial size evolution over the past several billion years. A comparison between the observed and model-predicted amount of size evolution will help to identify the mechanism(s) that are responsible. In Fig.~\\ref{M_dens} we directly compare the observed evolution in surface density with the predictions from the semianalytic work by \\citet{khochfar06b}. For galaxies with a given mass the model significantly under-predicts the evolution in size and density, except, perhaps, for the most massive galaxies. In our data set we see no indication that the magnitude of size and density evolution increases with galaxy mass, as predicted by the models. In fact, the most massive galaxies in our sample are precisely the only ones that are not different from local massive galaxies. Note, however, that statistically speaking the evidence for mass-dependent evolution is weak (see \\S~\\ref{results:mrevol}). Moreover, the most massive galaxies in our distant sample are a special subset, brightest cluster galaxies. Such galaxies have been shown to have properties that deviate from those of other massive galaxies \\citep[see, e.g.,][]{vonderlinden07, bernardi07b}. By including the $z=1.5-2.5$ photometric samples discussed in \\S~\\ref{dis:1} we can place further constraints on the models. In Fig.~\\ref{lit_M} we compare the observed size evolution of the available samples, normalized to $z=1$, with the model predictions from \\citet{khochfar06b}. Representing the model predictions by a single line is justified by the fact that the predicted evolution of $\\log(\\reff)$ with $\\log(1+z)$ is very close to linear. Again, the observed size evolution is stronger than that predicted by the model. It is interesting to note that, in qualitative agreement with the model prediction, we see a hint that size evolution depends on mass in the compilation presented in Fig.~\\ref{lit_M}. The samples containing on average the lowest-mass galaxies display marginally less evolution. It has to be kept in mind, however, that small sample sizes and systematic effects are more important for determining second-order effects such as mass dependence (see \\S~\\ref{results:mrevol}). A clue that systematic uncertainties may play a role is the remaining difference between the kinematic and photometric samples. Alternatively, it may signify non-linear evolution of $\\log(\\reff)$ with $\\log(1+z)$. \\begin{figure}[t] \\epsscale{1.2} \\plotone{lit_z.eps} \\caption{Size evolution with redshift as derived in this paper with dynamically determined masses (\\textit{large filled circles}) compared with previous results based on photometric masses (\\textit{small filled circles}). The solid line connects our samples at $z\\sim 0.06$ and $z\\sim 1$, the dashed line is a linear least-squares fit to the small filled data points. The open circles are samples of cluster galaxies with photometrically measured masses and serve as an illustration that size evolution shows a continuous trend between $z=2.5$ and the present. The broad agreement in size evolution as derived from galaxies with dynamically and photometrically determines masses reinforces the conclusions of previous, photometric studies whose results were potentially mitigated by considerable systematic effects that do not affect our analysis. } \\label{lit_z} \\end{figure} \\begin{figure}[t] \\epsscale{1.2} \\plotone{lit_M.eps} \\caption{Size evolution per unit redshift vs. mean galaxy mass of our sample (\\textit{large circle}) and samples taken from the literature (small squares; see Fig.~\\ref{lit_z} for references). Samples consisting of high-mass galaxies show somewhat stronger size evolution than samples consisting of low-mass galaxies, which is qualitatively, but not quantitatively, consistent with the predictions from the semianalytic model from \\citet{khochfar06b} (\\textit{solid line}). This conclusion should be considered highly tentative, however, as this interpretation is hampered by systematic uncertainties and small sample sizes. } \\label{lit_M} \\end{figure} \\subsection{Size Evolution of Individual Galaxies}\\label{dis:3} It appears that the observed size evolution of a factor of $\\sim$2 between $z=1$ and the present for early-type galaxies with masses $\\sim10^{11}~\\msol$ is similar to the predicted evolution for early-type galaxies that are an order of magnitude more massive (see Figs.~\\ref{M_dens} and \\ref{lit_M}). This suggests that the mechanism responsible for increasing the average size of early-type galaxies with time may be well understood, but that it is not implemented correctly in the current semianalytic model from \\citet{khochfar06b}. The process of size evolution may occur at different times and under different circumstances than is now assumed. This may be related to the late assembly of very massive galaxies in models of this kind \\citep[see also, e.g.,][]{delucia06}, a prediction that is challenged by various observations \\citep[e.g.,][]{cimatti06, scarlata07, cool08}. It is beyond the scope of this paper to fully discuss these possible discrepancies. Instead we will explore the question whether the proposed physical processes responsible for size evolution are consistent with the observed trends. In the semi-analytic models it is assumed (and this is confirmed by numerical simulations) that mergers drive size evolution. The gas content of merging galaxies largely determines the relative size of the merger remnant compared to its ancestors. Because gas fractions were higher in the past, galaxies that form early will be smaller than galaxies that form late. In the framework of cosmological simulations this means that galaxies at high redshift will be smaller because they were formed through gas-rich mergers and that those merger remnants can grow over time through subsequent mergers with other galaxies that are progressively more devoid of gas. The question is whether the observed size evolution is dominated by size evolution of individual galaxies or simply by the addition of larger galaxies over time. At $z=1$ only about 30--50\\% of the present-day early-type galaxy evolution had formed \\citep{bell04b, brown07, scarlata07, faber07}. If we assume that these galaxies will make up the 30--50\\% most dense early-type galaxies in the present-day universe, then the scatter in the local $\\mvir$-$\\reff$ relation implies that the sample-averaged size increases by a factor of 1.3--1.4 between $z=1$ and the present. Such evolution is thus expected in the absence of size evolution of individual galaxies and this is less than the observed evolution of a factor of 2. To explain the observed evolution by growth of the early-type galaxy population without changes in the sizes of individual galaxies, the number density of early-type galaxies is required to increase by an order of magnitude between $z=1$ and the present. Such strong evolution is clearly ruled out by the above-mentioned determinations of the number density of red galaxies at $z\\sim 1$. Similarly, at $z=2$ only $\\sim$10\\% of the galaxies with masses $\\gtrsim~10^{11}~\\msol$ had been assembled \\citep{kriek08b}; if those galaxies, evolve into the 10\\% most dense present-day early-type galaxies then an increase in average size by a factor of $\\sim$2 can be accounted for, less than the observed amount of evolution. These arguments are in agreement with the conclusions from \\citet{cimatti08}, who show that local galaxies with the same sizes and masses as galaxies in the $z=1-2$ samples are so rare in the local universe that it can be confidently ruled out that their structure remains unchanged up until the present day. Note, however, that these arguments may be affected by the aforementioned biases in the SDSS (\\S~\\ref{nearby:sigrad}). We conclude that size evolution due to the addition of larger galaxies over time contributes at most half of the observed evolution in the $\\mvir$-$\\reff$ relation. The remainder must be due to size evolution of individual galaxies. Numerical simulations have demonstrated that when early-type galaxies accrete neighbors without significant dissipational processes $\\sige$ does not change by much and that, to first order, $\\reff$ increases linearly with mass. This does not depend strongly on the mass of the accreted object, i.e., the mass ratio of the merger \\citep{boylan05, robertson06, boylan06}. Simulations in a cosmological context show that an increase in size by a factor of 2 between $z\\sim 1$ and the present is certainly possible \\citep{naab07}. The strong observed size evolution thus argues in favor of a scenario in which significant mass from low-mass companions is accreted onto existing early-type galaxies over the past $\\sim$7 Gyr, which also explains the broad tidal features that are frequently observed around early-type galaxies \\citep{vandokkum05}. As shown by \\citet{feldmann08} such features are not necessarily, and are even quite unlikely to be, the result of major merger events and are most likely due to the accretion of low-mass, gas-poor satellites. We note that the size evolution of individual galaxies and the evolution of the sample average are inseparable because galaxies evolve in mass as well as in size. Nonetheless, it is important to distinguish this complex scenario from the simple picture in which early-type galaxies that form at different redshifts have different sizes but do not structurally evolve at later times. The strong observed size evolution clearly rules out the latter, indicating that the build-up of the early-type galaxy population is a complex and ongoing process. Finally, it is remarkable that the change in the sizes of early-type galaxies is consistent with and differs by less than 15\\% from the change in the scale factor of the universe, $1+z$. Within the standard cold dark matter scenario this is likely a coincidence since dissipational, strongly non-linear processes that are decoupled from cosmic expansion dominate at the kiloparsec scale of forming galaxies. Nonetheless, we cannot exclude the possibility that there is an underlying, fundamental reason that galaxies are scale-invariant with respect to a co-moving coordinate system. In an alternative description of dark matter, i.e., Bose-Einstein condensed, ultra-light particles with a $\\sim$10 kpc-sized wave function \\citep[fuzzy dark matter or FDM,][]{sin94,hu00}, sizes of halos and their occupying galaxies possibly follow the cosmic expansion rate \\citep{lee08}." }, "0808/0808.4132_arXiv.txt": { "abstract": "We propose an explanation to the puzzling appearance of a wide blue absorption wing in the He~I~$\\lambda10830 \\rm{\\AA}$ P Cygni profile of the massive binary star $\\eta$ Car several months before periastron passage. Our basic assumption is that the colliding winds region is responsible for the blue wing absorption. By fitting observations, we find that the maximum outflow velocity of this absorbing material is $\\sim 2300 \\km \\s^{-1}$. We also assume that the secondary star is toward the observer at periastron passage. With a toy-model we achieve two significant results. (1) We show that the semimajor axis orientation we use can account for the appearance and evolution of the wide blue wing under our basic assumption. (2) We predict that the Doppler shift (the edge of the absorption profile) will reach a maximum $0-3~{\\rm weeks}$ before periastron passage, and not necessarily exactly at periastron passage or after periastron passage. ", "introduction": "\\label{sec:intro} There is no boring feature when it comes to the massive binary system \\astrobj{$\\eta$ Car}. Its $P=5.54 \\yr$ ($P =2022.7 \\pm 1.3~$d; Damineli et al. 2008a) periodicity has been observed in the radio (Duncan \\& White 2003), IR (Whitelock et al. 2004), visible (e.g., van Genderen et al. 2006), X-ray (Corcoran 2005), and in many emission and absorption lines (e.g., Damineli et al. 2008a,b). Every orbital cycle the system experiences a spectroscopic event, defined by the fading, or even disappearance, of high-ionization emission lines (e.g., Damineli 1996; Damineli et al. 1998, 2000, 2008a,b; Zanella et al. 1984). The rapid changes in the continuum, lines, and in the X-ray properties (e.g., Martin et al. 2006,a,b; Davidson et al. 2005; Nielsen et al. 2007; van Genderen et al. 2006; Damineli et al. 2008b; Corcoran 2005) are assumed to occur near the periastron passages of the highly eccentric, $e \\simeq 0.9$, binary orbit (e.g., Hillier et al. 2006). The He~I~$\\lambda10830 \\rm{\\AA}$ high excitation line has a complex P Cygni profile, composed of three blue-shifted peaks with significant variations over the cycle (Damineli et al. 1998; 2008b). The emission profile has significant variations over the cycle. The Doppler shifts of the peaks are of relatively low velocities, $\\vert v_{\\rm peaks} \\vert < 300 \\km \\s^{-1}$ (Damineli et al. 2008b). The location of the minimum of the profile (the deepest point in absorption) does not change with orbital phase, and stays at $v_{obs-m}=-570 \\km \\s^{-1}$. However, the absorption profile does change. In particular, just before periastron a wide blue wing appear in absorption, reaching $\\sim -1000 \\km \\s^{-1}$ a month before periastron, and $\\sim -1800 \\km \\s^{-1}$ at periastron, and the maximum equivalent width of absorption occurs $10 \\days$ after periastron passage (Damineli et al. 2008b). Damineli et al. (2008b) approximated the average radial velocity of the absorption profile at half intensity, and found it to change from $-640 \\km \\s^{-1}$ before phase zero to $-450 \\km \\s^{-1}$ shortly after phase zero. An additional He~I~$\\lambda10830 \\rm{\\AA}$ blue absorption feature of up to $-1000 \\km \\s^{-1}$, is observed at several arcseconds from the center in the lobes (Smith 2002). In this paper we refer only to the He~I~$\\lambda10830 \\rm{\\AA}$ ground measurements of Damineli et al. (2008b). We note that other absorption and emission lines of He~I can be formed in different regions in the binary system (see also Kashi \\& Soker 2007b). In particular, some visible He~I lines can be formed in the hot winds of the two stars close to their origin, as compared with the He~I~$\\lambda10830 \\rm{\\AA}$ that is formed in cooler regions. We show that this cooler region can be the post-shock primary wind. Therefore, different He~I lines need not have the same behavior along the orbit. Because of the winds' very complicated flow structure, when starting this project we limited ourself to build a toy-model in order to achieve two goals: (1) To show that the orientation where the secondary is toward us at periastron ($\\omega = 90^\\circ$), can be accounted for the development of a wide blue absorption wing starting several weeks before periastron passage. (2) To encourage a nightly observation of the He~I~$\\lambda10830 \\rm{\\AA}$ close to periastron passage. ", "conclusions": "\\label{sec:es} We study the blue absorption wing of the He~I~$\\lambda10830 \\rm{\\AA}$ P Cygni profile of $\\eta$ Car. The two winds of the two stars collide, and the post shocked gas of the two winds flow on both sides of a surface (the discontinuity surface) that has a pseudo-conical structure: the conical shell. The shocked secondary's wind accelerates part of the shocked primary's wind to high velocity (Figure \\ref{fig:colliding}). This gas, and the segments of the shocked secondary's wind which cool to a low temperature near the contact discontinuity (Figure \\ref{fig:colliding}), are assumed to be responsible for the blue absorption wing. This is the fundamental assumption of our model. We use our previous results and assume an orientation where the secondary is toward the observer at periastron ($\\omega=90^\\circ$; see Figure \\ref{fig:omega}). For the conical shell we built a toy model (Figure \\ref{fig:cone}). The absorption profile, up to a scaling factor, is calculated according to equations (\\ref{eq:sum}) and (\\ref{eq:I}), and the results are presented in Figure \\ref{fig:abs}. We are interested in the bluest part of the absorption profile, where intensity is lower; in our scaled units these are the contours in the range $I \\simeq 0.8-1$. Using our fundamental assumption, and a toy model for the conical shell of the colliding winds, we showed that with our orbital orientation we can account for the appearance of the wide blue wing several months before periastron. Other semimajor orientations $\\omega$, cannot reproduce the results under our fundamental assumption (figure \\ref{fig:otheromega}) This is our main result, namely, that if the absorber responsible for the blue wing of the He~I~$\\lambda10830 \\rm{\\AA}$ line reside in the winds collision region, then only the $\\omega \\simeq 90^\\circ$ can account for the blue wing of this line. Our results also predict that the Doppler shift $v_D$ (the edge of the profile) will reach a maximum $0-3~\\rm{weeks}$ before periastron passage. Since close to periastron the conical shell starts to collapse onto the primary, and even before it experiences the effect of wrapping, it is hard to pinpoint the exact time of this maximum. Nevertheless this maximum should be observed before the event." }, "0808/0808.1521_arXiv.txt": { "abstract": "This is the first of a series of papers devoted to the investigation of a large sample of brightest cluster galaxies (BCGs), their kinematic and stellar population properties, and the relationships between those and the properties of the cluster. We have obtained high signal-to-noise ratio, long-slit spectra of these galaxies with Gemini and WHT with the primary purpose of investigating their stellar population properties. This paper describes the selection methods and criteria used to compile a new sample of galaxies, concentrating on BCGs previously classified as containing a halo (cD galaxies), together with the observations and data reduction. Here, we present the full sample of galaxies, and the measurement and interpretation of the radial velocity and velocity dispersion profiles of 41 BCGs. We find clear rotation curves for a number of these giant galaxies. In particular, we find rapid rotation ($>$ 100 km s$^{-1}$) for two BCGs, NGC6034 and NGC7768, indicating that it is unlikely that they formed through dissipationless mergers. Velocity substructure in the form of kinematically decoupled cores is detected in 12 galaxies, and we find five galaxies with velocity dispersion increasing with radius. The amount of rotation, the velocity substructure and the position of BCGs on the anisotropy-luminosity diagram are very similar to those of ``ordinary'' giant ellipticals in high density environments. ", "introduction": "The galaxies in the centres of clusters are unique. They are usually the dominant, brightest and most massive, galaxies in their clusters. A wealth of imaging data has been accumulated for these intriguing objects (Malumuth $\\&$ Kirshner 1985; Schombert 1986, 1987, 1988; Postman $\\&$ Lauer 1995; Collins $\\&$ Mann 1998; Brough et al.\\ 2002; Laine et al.\\ 2003), however the spectroscopic data is limited to very small samples or narrow wavelength coverage. Tonry (1984), Tonry (1985), Gorgas, Efstathiou $\\&$ Aragon-Salamanca (1990), Fisher, Illingworth $\\&$ Franx (1995a), Fisher, Franx $\\&$ Illingworth (1995b), Cardiel, Gorgas $\\&$ Aragon-Salamanca (1998) and Carter, Bridges $\\&$ Hau (1999) each investigated 18 or less galaxies, and moreover measured only three or less indices. Brough et al.\\ (2007) used a wide wavelength range but included only three brightest cluster galaxies (BCG). Von der Linden et al.\\ (2007) studied BCGs from the Sloan Digital Sky Survey (SDSS), however the spatial information is lacking since fibers were used. Among BCGs there exists a special class of galaxies catalogued as cD galaxies, although there exists a great deal of confusion over the exact meaning of the classifications gE, D and cD. Matthews, Morgan $\\&$ Schmidt (1964) outlined the following definition: ``D galaxies have an elliptical-like nucleus surrounded by an extensive envelope. The supergiant D galaxies observed near the centre of a number of Abell's rich clusters have diameters 3 -- 4 times as great as the ordinary lenticulars in the same clusters. These very large D galaxies observed in clusters are given the prefix ``c'', in a manner similar to the notation for supergiant stars in stellar spectroscopy''. Later, Schombert (1987) defined gE galaxies as distinct from other early-type galaxies by their large size; D galaxies as being gE galaxies with a shallow surface brightness profile slope; and cD galaxies as D galaxies with large extended stellar haloes. The last are also more diffuse than normal ellipticals (Schombert 1986). Because of the confusion surrounding the definition of D galaxies, and the fact that they are rarely regarded as a separate type of object in the modern literature, this study will only refer to cD and non-cD BCGs (i.e BCGs containing a halo or not). For BCGs, we adopt the definition to comply with recent literature (for example Von der Linden et al.\\ 2007), where BCG refers to the central, dominant galaxy in a cluster. For a small fraction of clusters, the BCG might not strictly be the brightest galaxy in the cluster. Approximately 20 per cent of rich clusters of galaxies contain a dominant central cD galaxy (Dressler 1984; Oegerle $\\&$ Hill 2001), although they can be found in poor clusters as well (Giacintucci et al.\\ 2007). Some clusters have more than one cD galaxy, but a cD galaxy is always the dominant member of a local subcluster. The surface-brightness profiles of cD galaxies are displaced above the de Vaucouleurs law (de Vaucouleurs 1948) at large radii. The break in the cD galaxy surface-brightness profile typically occurs between 24 and 26 mag/arcsec$^{2}$ in the V-band (Sarazin 1988). The interpretation of this deviation is that the galaxy is embedded in an extensive luminous stellar halo. Three main theories have been proposed over the last four decades to explain the properties of BCG, and in particular cD, galaxies. \\paragraph*{Theory 1.} In the first theory, BCG formation is caused by the presence of \\textit{cooling flows} in clusters of galaxies (Cowie $\\&$ Binney 1977). Cooling flow clusters are common in the local universe and BCGs are most often found at the centres of these systems (Edwards et al.\\ 2007). If the central cluster density is high enough, intracluster gas can condense and form stars at the bottom of the potential well. Observations that support this idea are blue- and UV-colour excesses observed in the central galaxy of Abell 1795 (indicative of star formation) by McNamara et al.\\ (1996) and molecular gas detected in ten out of 32 central cluster galaxies by Salom\\'{e} $\\&$ Combes (2003). Cardiel et al.\\ (1998) obtained radial gradients for the D$_{4000}$ and Mg$_{2}$ spectral features in 11 central cluster galaxies. Their observations were consistent with an evolutionary sequence in which radio-triggered star formation bursts take place several times during the lifetime of the cooling flow in the centre of the cluster. However, McNamara $\\&$ O'Connell (1992) find only small colour anomalies with small amplitudes, implying star formation rates that account for at most a few percent of the material that is cooling and accreting onto the central galaxy. More recently, \\textit{XMM-Newton} observations showed that the X-ray gas in cluster centres does not cool significantly below a threshold temperature of $kT\\sim1-2$ keV (Jord\\'an et al.\\ 2004, and references therein). The central cluster galaxies often host radio-loud AGN which may account for the necessary heating to counteract radiative cooling (von der Linden et al.\\ 2007). Although BCGs are probably not completely formed in cooling flows, the flows play an important role in regulating the rate at which gas cools at the centres of groups and clusters. \\paragraph*{Theory 2.} The second theory was proposed by Merritt (1983) and suggests that the essential properties of BCGs, and in particular those with haloes, are determined when the clusters collapse (\\textit{primordial origin}). Thereafter, frequent mergers of galaxies would be inhibited by the relatively high velocities between galaxies. Merritt (1983) argued that all galaxies had large haloes early in the life of the cluster. These haloes were then removed by the mean cluster tidal field during the initial collapse and returned to the cluster potential, except for the central member which remained unaffected because of its special position with respect to the cluster potential. \\paragraph*{Theory 3.} The third, and most widely accepted, theory is in the context of the $\\Lambda$CDM cosmology and relates the formation of the central galaxy to mergings with or captures of less massive galaxies, and is known as \\textit{``galactic cannibalism\"}. It was first proposed by Ostriker $\\&$ Tremaine (1975) and developed by Ostriker $\\&$ Hausman (1977). The most complete quantitative prediction of the formation of BCGs in the now standard CDM model of structure formation is by De Lucia $\\&$ Blaizot (2007). They used N-body and semi-analytic techniques to study the formation and evolution of BCGs and found that, in a model where cooling flows are suppressed at late times by AGN activity, the stars of BCGs are formed very early (50 per cent at $z \\sim 5$ and 80 per cent at $z \\sim 3$) and in many small galaxies. They also found that BCGs assemble late: half of their final mass is typically locked up in a single galaxy after $z \\sim 0.5$ (illustrated in their figure 9). A very similar conclusion was reached by Romeo et al.\\ (2008), who performed N-body and hydrodynamical simulations of the formation and evolution of galaxy groups and clusters in a $\\Lambda$CDM cosmology to follow the build-up of two clusters and 12 groups. Observationally, Aragon-Salamanca et al. (1998) examined the K-band Hubble diagram for BCGs up to a redshift of $z$ = 1. They found that the BCGs had grown by a factor of two to four since $z$ = 1. Brough et al. (2002) found a similar result but discovered that the mass growth depended on the X-ray luminosity of the host cluster. They found that BCGs in high X-ray luminosity clusters showed no mass accretion since $z$ = 1 as opposed to BCGs in low X-ray luminosity clusters which grew by a factor of four. However, the recent near-infrared photometric study of 42 BCGs by Whiley et al.\\ (2008) in the 0.2 $<$ $z$ $<$ 1 range contradicts this. They studied the colour and rest-frame $K$-band luminosity evolution of BCGs and found it to be in good agreement with population synthesis models of stellar populations which formed at $z \\sim $ 2 and evolved passively thereafter. Using the Millennium Simulation, De Lucia et al.\\ (2006) studied how formation histories, ages and metallicities of elliptical galaxies depend on environment and on stellar mass. Their Figure 9 shows the effective number of progenitors of early-type galaxies as a function of galaxy stellar mass. The number of effective progenitors is less than two for galaxies up to stellar masses of $\\simeq10^{11}$ M$_{\\odot}$. This function suddenly increases up to the value of approximately five effective progenitors for the mass of a typical BCG. A related theory, called \\textit{tidal stripping}, was first proposed by Gallagher $\\&$ Ostriker (1972). Cluster galaxies that pass near the gravitational centre of the cluster may be stripped of some of their material by the tidal forces from the cluster potential or the central galaxy potential. The stripped material falls to the centre of the potential well, and could contribute to the observed haloes of cD galaxies. The most massive galaxies surrounding the central galaxy would be preferentially depleted as they are most strongly affected by dynamical friction (Jord\\'an et al.\\ 2004). The difference between stripping and primordial origin is that stripping (and cD halo formation) begins after cluster collapse whereas primordial origin assumes that the tidal events occur before collapse, and that the cD halo is not a consequence of tidal stripping (Schombert 1988). The observation of multiple nuclei in central galaxies favours the cannibalism theory (Postman $\\&$ Lauer 1995). Tonry (1984) observed the velocity and velocity dispersion profiles of NGC6166 and NGC7720 and their multiple nuclei. He found that the stellar velocity dispersion of the central galaxy demonstrated that the multiple nuclei are not following circular orbits. This was followed by a bigger sample of 14 multiple nuclei BCGs for which the redshifts and stellar velocity dispersions are presented in Tonry (1985). Yamada et al.\\ (2002) showed that the BCG in a cluster at $z=1.26$ is composed of two distinct sub-units that are likely to fully merge on a time-scale of $10^{8}$ years. Jord\\'an et al.\\ (2004) studied the globular cluster systems in BCGs with confirmed haloes from Hubble Space Telescope observations. They concluded that the observed globular cluster metallicity distributions are consistent with those expected if BCGs galaxies form through cannibalism of numerous galaxies and protogalactic fragments that formed their stars and globular clusters before capture and disruption, although they state that the cannibalism scenario is not the only possible mechanism to explain these observations. The Jord\\'an et al.\\ (2004) globular cluster data also suggests that BCGs experienced their mergers prior to cluster virialisation, yet the presence of tidal streams suggest otherwise (Seigar, Graham $\\&$ Jerjen 2007). Carter $\\&$ Metcalfe (1980) and West (1989) showed that the major axis of BCGs tends to be aligned with the major axis of the cluster galaxy distribution. Recent studies of the Coma cluster (Torlina, De Propris $\\&$ West 2007) show strong evidence that there are no other large-scale galaxy alignments other than for the BCGs. This also confirms that BCGs form via mechanism related to collimated infall of galaxies along the filaments and the growth of the cluster from the surrounding large scale structure (Boylan-Kolchin, Ma $\\&$ Quataert 2006). In the CDM cosmology it is now understood that local massive cluster galaxies assemble late through the merging of smaller systems. In this picture, cooling flows are the main fuel for galaxy mass-growth at high redshift. This source is removed only at low redshifts in group or cluster environments, due to AGN feedback (De Lucia $\\&$ Blaizot 2007). Some of the outstanding issues are: whether BCGs have a different formation mechanism than elliptical galaxies; and if the formation of BCGs are controlled by environment. To address these issues the present project was initiated, and long-slit spectra of a large and statistically significant sample of BCGs were obtained. These data provide a set of spectral indices, covering a wide wavelength range. The luminous central galaxies will be contrasted with other early-type galaxies to look for relative differences in their evolution, using features sensitive to stellar population age and abundances of various elements. The first part of this project entails the determination of the stellar kinematics through derivation of velocity and velocity dispersion profiles. A forthcoming paper will be devoted to the measurement and analysis of line strengths for this sample of BCGs. This paper is organised as follows. Section 2 details the sample and the selection criteria, followed by a description of the observations and data reductions in Section 3. The kinematical measurements are described in Section 4 and the individual galaxy kinematic profiles and notes are presented in Section 5. Section 6 discusses the kinematic properties of BCGs as a class, compared to other Hubble types. Conclusions and future work are given in Section 7. ", "conclusions": "\\label{kinematicanalysis} (i) \\textit{Five out of 41 BCGs (ESO349-010, ESO444-046, ESO552-020, NGC3311, PGC026269) were found to have a positive velocity dispersion gradient.} The radial kinematic studies done so far on early-type galaxies are mostly limited to normal ellipticals, for which flat or decreasing velocity dispersion profiles are found. The majority of the results previously obtained for very small samples of BCGs are similar to those of normal ellipticals. Fisher et al.\\ (1995a) found one galaxy (IC1101) in their sample of 13 BCGs with a positive velocity dispersion gradient. Carter et al.\\ (1999) found one (NGC6166) of their sample of three BCGs to have a positive velocity dispersion gradient, although Fisher et al.\\ (1995a) did not find it for this galaxy. Brough et al.\\ (2007) found negative velocity dispersion gradients in five out of their sample of six brightest cluster and group galaxies (the other one had a zero velocity dispersion gradient). Both IC1101 and NGC6166 form part of the sample of BCGs studied here, but in both cases the measured velocity dispersion profiles do not reach the radius achieved in the much smaller samples in the above mentioned studies. Thus, the positive velocity dispersion gradient could not be confirmed for IC1101 or NGC6166 (see Figures \\ref{fig:NGC6166kin} and \\ref{fig:IC1101kin}). However, five other BCGs were found to have a positive velocity dispersion gradient, though admittedly the slope is marginally positive in some cases. If these positive velocity dispersion gradients are not caused by systematic errors in the various data reduction or velocity dispersion measurements by different authors, then they imply a rising mass-to-light ratio. \\medskip (ii) \\textit{At least 12 BCGs (NGC3842, NGC4889, NGC7647, ESO488-027, IC5358, MCG-02-12-039, NGC1713, NGC2832, NGC4839, NGC6269, NGC7649 and UGC05515) show clear velocity substructure in their profiles.} From studies of elliptical galaxies in high density environments (e.g. Koprolin $\\&$ Zeilinger 2000), the incidence of KDCs is observed to be about 33 per cent, rising to 50 per cent when projection effects are considered. Hau $\\&$ Forbes (2006) finds KDCs in 40 per cent of their isolated elliptical galaxies. For BCGs, we have found at least 12 of the 41 BCGs show clear velocity substructure, amounting to 29 per cent of the sample (intermediate and minor axis data included). KDCs can be the result of a merger event (Koprolin $\\&$ Zeilinger 2000), but can also occur when the galaxy is triaxial and supports different orbital types in the core and main body (Statler 1991). The fact that the incidence of KDCs in BCGs compares with that found for normal elliptical galaxies in high density environments suggests that the two classes share the same fraction of galaxies with triaxial shapes. \\medskip (iii) \\textit{NGC6034 and NGC7768 possess significant rotation (134 and 114 km s$^{-1}$ respectively) along the major axis. Several other BCGs show rotation that is $>$ 40 km s$^{-1}$ and more than three times the standard error: ESO346-003, GSC555700266 and NGC4839 (major axis spectra); ESO488-027, IC5358 and UGC05515 (intermediate axis spectra), as do the two elliptical galaxies NGC4946 (major axis spectra) and NGC6047 (intermediate axis spectra).} Carter et al.\\ (1999) found small rotation along the major axis at large radii (30 -- 40 arcsec) for their sample of three BCGs, which is consistent with the nearly complete lack of rotation found near the centres of the sample of 13 BCGs by Fisher et al.\\ (1995a). According to Fisher at al.\\ (1995a), the lack of rotation found in samples of BCGs are in agreement with the expectation of declining importance of rotation with increasing luminosity for elliptical galaxies. The lack of rotation is also compatible with the idea that these objects formed through dissipationless mergers (Boylan-Kolchin et al.\\ 2006). The remnants left by mergers with or without dissipation are expected to differ in their kinematical structure. In a merger which involves gas-rich galaxies, the gas will form a disk. After the gas has been removed from the system at the end of the merger (through ejection and converted into stars), the remnant will show rotation (Bournaud et al. 2005). Whereas in a merger where dissipationless processes dominate, the remnant will show little or no rotation (Naab $\\&$ Burkert 2003; Cox et al. 2006). In this study, clear rotation above 100 km s$^{-1}$ was found for NGC6034 and NGC7768, while most BCGs showed little or no rotation. This kinematical differentiation (the existence of slow and fast rotators) in early-type galaxies is also clearly visible in the SAURON data presented by Emsellem et al.\\ (2006). The amount of flattening that is expected due to rotation in a galaxy depends on the balance between ordered and random motions, and this can be quantified using the anisotropy parameter. The rotation of elliptical galaxies is conventionally expressed as the anisotropy parameter, defined as $(V_{\\rm max}/\\sigma_{0})^{\\ast} =$ ($V_{\\rm max}/\\sigma_{0}$)/$\\sqrt{\\epsilon/1-\\epsilon}$\\ \\ (Kormendy 1982), where the rotational velocity $V_{\\rm max}$ is measured as in Section \\ref{Sec:kinematics}, and the central velocity dispersion $\\sigma_{0}$ is taken as the measurement for the central velocity dispersion in Table \\ref{table:GalVel}. A value of $(V_{\\rm max}/\\sigma_{0})^{\\ast} \\approx 1$ would be expected if a galaxy is flattened by rotation. The anisotropy parameter $(V_{\\rm max}/\\sigma_{0})^{\\ast}$ can be used to separate galaxies that are rotationally supported from those that are supported by $\\sigma$ anisotropy where the value is substantially less than unity. The division occurs at about $(V_{\\rm max}/\\sigma_{0})^{\\ast} = 0.7$ (Bender, Burstein $\\&$ Faber 1992). Figure \\ref{fig:Anisotropy} shows the anisotropy parameter as a function of galaxy $B$-band luminosity, where the above mentioned division is indicated by the horizontal dashed line. Note that the errors indicated are only propagated from the errors on the velocity measurements taken to be the extreme radial velocity points, and the errors on the central velocity dispersion. They do not take into account the general uncertainties involved in determining the most extreme velocity measurements. Therefore, care has to be taken when interpreting individual points on the diagram. The only notable BCG data point that lies significantly above the dashed line is for the galaxy PGC026269, which possesses moderate rotation (51 km s$^{-1}$) and surprisingly low central velocity dispersion (222 km s$^{-1}$). However, the ellipticity of this galaxy is zero and it is not rotationally supported. Another factor that complicates the dynamical interpretation of individual points is that the observed ellipticity is a global property of the galaxy, whereas the kinematic measurements taken here only reflect the kinematics along the axis where the slit was placed, and only close to the centre of the galaxy. For example, a disk component may dominate the measured kinematics but will have little effect on the ellipticity, making the galaxy appear to rotate faster than its global ellipticity would suggest (Merrifield 2004). For comparison, the sample of isolated ellipticals from Hau $\\&$ Forbes (2006) are also plotted in Figure \\ref{fig:Anisotropy}. Of this sample, 11 galaxies were observed along the major axis, and the central velocity dispersions were derived from the bins closest to the galaxy cores. The BCGs show less rotational support than the isolated elliptical galaxies, as a class. Spiral bulge (typically rotationally supported) and giant elliptical data from Bender et al.\\ (1992) are also plotted. Their central velocity dispersions were derived over the whole half-light radii of the galaxies. Even though large rotation velocities were found for a few individual cases, the BCGs are consistent with the general trend for very massive galaxies. \\begin{figure} \\centering \\includegraphics[scale=0.32]{Anisotropy.eps} \\caption[The Anisotropy-luminosity Diagram]{The anisotropy-luminosity diagram. The horizontal dashed line separates the rotationally supported galaxies from the anisotropic galaxies as described in the text. Only the BCGs for which major axis spectra were taken (within 10 degrees) are plotted. The box plotted in the figure outlines the region containing data on giant ellipticals from Bender et al.\\ (1992).} \\label{fig:Anisotropy} \\end{figure}" }, "0808/0808.0461_arXiv.txt": { "abstract": "The 9-month SWIFT Burst Alert Telescope (BAT) catalog provides the first unbiased ($N_H < 10^{24}$\\,cm$^{-2}$) look at local ($ = 0.03$) AGN. In this paper, we present the collected X-ray properties (0.3 -- 12\\,keV) for the 153 AGN detected. In addition, we examine the X-ray properties for a complete sample of non-beamed sources, above the Galactic plane (b$\\ge 15^{\\circ}$). Of these, 45\\% are best fit by simple power law models while 55\\% require the more complex partial covering model. One of our goals was to determine the fraction of ``hidden'' AGN, which we define as sources with scattering fractions $\\le 0.03$ and ratios of soft to hard X-ray flux $\\le 0.04$. We found that ``hidden'' AGN constitute a high percentage of the sample (24\\%), proving that they are a very significant portion of local AGN. Further, we find that the fraction of absorbed sources does increase at lower unabsorbed 2--10\\,keV luminosities, as well as accretion rates. This suggests that the unified model requires modification to include luminosity dependence, as suggested by models such as the 'receding torus' model (Lawrence 1991). Some of the most interesting results for the BAT AGN sample involve the host galaxy properties. We found that 33\\% are hosted in peculiar/irregular galaxies and only 5/74 hosted in ellipticals. Further, 54\\% are hosted in interacting/merger galaxies. Finally, we present both the average X-ray spectrum (0.1--10\\,keV) and $\\log N$-$\\log S$ in the 2-10\\,keV band. With our average spectrum, we have the remarkable result of reproducing the measured CXB X-ray power law slope of $\\Gamma \\approx 1.4$ (Marshall {\\it et al.} 1980). From the $\\log N$-$\\log S$ relationship, we show that we are complete to $\\log S \\ge -11$ in the 2--10\\,keV band. Below this value, we are missing as many as 3000 sources at $\\log S = -12$. Both the collected X-ray properties of our uniform sample and the $\\log N$-$\\log S$ relationship will now provide valuable input to X-ray background models for $z \\approx 0$. ", "introduction": "Active galactic nuclei (AGN) are among the most powerful sources of energy in the Universe, where the brightest quasars can outshine all of the stars in their host galaxy by 100 times. Optical studies of AGN reveal strong narrow and broad emission lines indicative of AGN activity in the nearby Universe (z$ < 1$). However, X-ray and optical surveys fail to select the same AGN samples \\citep{2004ASSL..308...53M}. X-ray surveys identify more sources whose 2 -- 10\\,keV emission is obscured by high column density absorbing material in the line of sight. Still, even the X-ray surveys are affected by heavy obscuration, making it difficult to detect sources with column densities $> 10^{24}$\\,cm$^{-2}$ (Compton thick sources). This provides a major question that AGN surveys need to address: how many heavily obscured or Compton thick sources exist? \\citet{2000MNRAS.318..173M} estimated that the number could be as high as an order of magnitude more than the unobscured sources, which are easily detected in optical and soft X-ray surveys. In order to account for these additional heavily obscured sources and determine their contribution to the cosmic X-ray background, very hard X-ray ($> 10$\\,keV) surveys are needed. Only at these wavelengths is the AGN emission penetrating enough to pass through much of the dust and gas in the line of sight. With the Burst Alert Telescope (BAT) on board SWIFT, we now have the first sensitive unbiased AGN survey, towards all but the most heavily obscured sources ($N_H > 10^{24}$\\,cm$^{-2}$). This is due to BAT's sensitivity at very high X-ray energies (14 -- 195\\,keV). With a much larger sample than previous (such as HEAO-1) and contemporary (Integral is most sensitive along the Galactic plane) very hard X-ray missions, BAT is the first sensitive all-sky very hard X-ray survey in 28 years. From the 9-month catalog of BAT AGN \\citep{2007arXiv0711.4130T}, a large enough sample has been obtained (153 sources) to determine the statistical properties. In the catalog paper, only the X-ray derived column density and a complexity flag were indicated. However, in this paper we provide a more detailed look at the X-ray properties, including archival data, analyses from the literature, and previously unanalyzed SWIFT X-ray Telescope (XRT) observations. This work follows upon an XMM-Newton follow-up study of 22 BAT AGN, in which we reported that the ``hidden''/buried AGN described in \\citet{2007ApJ...664L..79U} may be a significant fraction of the BAT AGN \\citep{2008ApJ...674..686W}. Our goals are two-fold. First, we present the X-ray properties for the entire 9-month catalog. Second, we examine in more detail the collective properties of a uniform sample. This sample consists of non-beamed sources with Galactic latitudes $\\ge 15^{\\circ}$. In Section~\\ref{bat-data}, we describe the observations and the spectral fits, including data from ASCA, XMM-Newton, Chandra, Suzaku, and SWIFT XRT. In Section~\\ref{bat-spectra}, we describe the general properties of the entire BAT 9-month AGN sample. In Section~\\ref{bat-unbiased}, we describe in more depth the properties of our uniform sample. In Section~\\ref{bat-cxb}, we present the average X-ray spectrum as well as the 2--10\\,keV $\\log N$-$\\log S$ relation. These X-ray properties can now be used as input to X-ray background models for $z \\approx 0$. We then discuss the properties of the host galaxies in Section~\\ref{bat-host}. Finally, we summarize our results in Section~\\ref{bat-summary}. ", "conclusions": "\\label{bat-summary} In this paper, we present the X-ray properties of a uniform sample of very hard X-ray (14 -- 195\\,keV) selected AGN. We present a number of interesting results that highlight the many uses of a uniform very hard X-ray survey. This paper is complimentary to the 9-month AGN survey paper \\citep{2007arXiv0711.4130T}, which presents the 14--195\\,keV properties of the sources. Additionally, this paper confirms the results of our earlier study on {\\it XMM-Newton} observations of a representative sample of the BAT AGN \\citep{2008ApJ...674..686W}. Among these, we show that: (1) the X-ray and optical classifications agree, i.e. Sy 1s have low X-ray column densities while Sy 2s are more obscured, (2) the average power law index, $\\Gamma \\approx 1.8$, agrees with the results from HEAO-1 \\citep{1982ApJ...256...92M}, (3) ``hidden'' AGN are a significant fraction of local AGN, where we can now quantify this value as $\\approx 24$\\%, and (4) nearly half (45\\%) of local AGN are well-fit by a simple model (all with $\\log N_H < 23$) while the remaining sources (55\\%) require a more complex model. In addition, this paper presents a number of additional, important results. From examining the host galaxy properties, we found that the majority of the X-ray obscuration is not simply from the host galaxy (by comparing host inclination ($b/a$) to X-ray column density). The most surprising results, however, were that many of the host galaxies are peculiar/irregular galaxies (33\\%). Further, an even larger fraction (54\\%) have either a close companion galaxy or are known mergers. This is observational proof that galaxy interactions may be driving activity in local supermassive black holes. However, we also find that the distribution of AGN 2--10\\,keV luminosities and accretion rates, as well as morphologies, are the same between interacting and non-interacting hosts. While, it is unclear what these results mean, however, there appears to be more than one trigger besides mergers for local AGN activity. Our team is currently compiling higher quality images and photometry to better quantify these results. From our uniform sample (102 sources with $|b| \\ge 15^{\\circ}$), we found that the distributions of both unabsorbed 2 -- 10\\,keV luminosity and accretion rate are significantly lower for Sy 2s than Sy 1s. While earlier studies found this connection in 2 -- 10\\,keV luminosity \\citep{2003ApJ...596L..23S, 2003ApJ...598..886U}, this is the first time it has been reported in accretion rate. We also showed that the fraction of obscured AGN is indeed larger for lower luminosities (absorption corrected $L_{2-10 keV}$) and accretion rates. However, we note that the most heavily obscured sources ($\\log N_H \\ge 23$) do not dominate this relationship. Since the unified model predicts differences between absorbed and unabsorbed sources are a product of viewing angle alone, our results provide a challenge, arguing in favor of a luminosity-dependent AGN model. Another result involves the correlation between accretion rate and $\\Gamma$. In \\citet{2008ApJ...674..686W}, we had found indication of a connection between Eddington ratio (or 2 -- 10\\,keV luminosity) and $\\Gamma$ using the spectral fits of multiple observations for individual sources. The fact that we did not observe a correlation in our larger sample seems to be a result of our sources having a larger range of Eddington ratios (or 2 -- 10\\,keV luminosities). We suggest that previous studies, for instance by Shemmer et al. 2006, see this correlation because their samples have a narrower range of properties (being mid- to high luminosity AGNs). The primary correlation appears to be with accretion rate and not hard band luminosity. Such a correlation should appear when comparing $\\Gamma$ to $L^{corr}_{2-10 keV}/L_{Edd}$ for multiple observations of individual sources or for a sample of sources with a narrow range of accretion rates. In a similar manner, we found that while our sample did not immediately confirm the X-ray Baldwin effect, binning the sources by luminosity, we were able to reproduce the anti-correlation between unabsorbed 2 -- 10\\,keV luminosity and Fe K EW. The primary anti-correlation, however, again appears to be with Eddington rate. When we binned the values by our Eddington ratio proxy, we found that $EW \\propto {L^{corr}_{2-10 keV}/L_{Edd}}^{0.26 \\pm 0.03}$ (agreeing with the results of \\citet{2007AA...467L..19B}). Since both $\\Gamma$ and Fe K EW are dependent on accretion rate, this suggests that the $\\Gamma$-EW correlation found by \\citet{2007ApJ...664..101M} is a result of the accretion rate dependences. Having classified the X-ray spectra of our sample into simple and complex categories, we were able to examine the properties of the two sub-samples in more detail. For the simple model sources, we found that 41\\% of the sources exhibited a soft excess. Having modeled this parameter with a simple blackbody model, we found the average temperature to be $kT = 0.10$\\,keV. We also found that there was a significant amount of scatter in this value ($\\sigma = 0.07$\\,keV), contrasting with the \\citet{2004MNRAS.349L...7G} results for PG quasars. We found no correlation between the blackbody temperature and Eddington ratio, black hole mass, or photon index. However, we did find a correlation between the luminosity of the blackbody component and the luminosity in the power law. This relationship is linear ($L_{pow} \\propto L_{kT}$) and may provide a challenge to the current soft excess models. Examining the complex model sources, we found that the majority of these sources included absorbed AGN. Of the 4 Sy 1s in this category, all have complex absorption features in their X-ray spectra. For these sources, we showed that the nature of the soft emission ($L_{0.5-2 keV}$) for these sources is unclear. Over half have soft band luminosities low enough to be the result of galactic emission from star formation/X-ray binaries. However, of the sources with higher soft luminosities, 3 are ``hidden''/buried AGN. This argues that the soft emission may be scattered AGN emission ($\\le 0.03$). An important result we found is that the ``hidden''/buried AGN, sources with a high covering fraction, are a significant fraction of local AGN. Among the complex sources, 45\\% are ``hidden''. For these sources, we found that the FIR luminosity is not consistent with an increased star formation rate, as suggested by \\citet{2007ApJ...664L..79U}. However, without higher quality X-ray spectra and multi-wavelength observations, we are unable to further explore the nature of these sources. While BAT is quite good at finding heavily obscured sources, we found that none of the 9\\,month sources in our uniform sample have spectra consistent with heavily obscured Compton-thick objects ($N_H > 1.4 \\times 10^{24}$\\,cm$^{-2}$). However, we do detect sources classified as Compton-thick in other studies based on a reflection dominated spectrum or strong Fe K EW (for instance 3C 452 and NGC 4945). Additionally, we did not include an analysis of the very complex source NGC 6240, which may also be classified as Compton thick. Since the Compton hump lies above 10\\,keV, spectral fits with and without reflection can be degenerate in the 0.1--10\\,keV band. Therefore, a full analysis of the Compton thick nature of the BAT sources must be deferred to future studies. One remarkable result we found came from the average spectrum we constructed in the 0.1--10\\,keV band with the measured spectral properties of our uniform sample. Here, our data reproduce the measured slope of the CXB ($\\approx 1.4$). This highlights the importance of the BAT survey in selecting heavily absorbed sources. More importantly, this is observational proof that the combination of BAT-detected absorbed and unabsorbed local AGN replicate the shape of the CXB. If the distribution of source properties at $z \\approx 1$, where much of the CXB originates, is the same as that of the BAT-detected AGN, the spectral paradox is resolved. To test our completeness in the 2--10\\,keV band, we plotted the distribution of $\\log N$-$\\log S$ for the entire uniform sample. This showed that while the sample is complete in the 14--195\\,keV band \\citep{2007arXiv0711.4130T}, we are only complete above $\\log S = -11$ in the 2--10\\,keV band. Further, this distribution suggests that we are missing as many as 3000 sources at $\\log S = -12$, requiring that these sources have 14--195\\,keV fluxes below the current flux limit of the BAT survey. Possibly these sources are ``hidden'' AGN with even higher X-ray columns ($\\log N_H \\ge 24$). Such sources must have a high ratio of $F_{2-10 keV}/F_{14-195 keV}$, like NGC 1068. Also, they may or may not contain Compton-thick sources, an answer to which our data can not supply. These results, in addition to the X-ray properties (including column densities and spectral indices) will provide important input for CXB models at low redshift ($z \\approx 0$). Overall, our analysis of the X-ray properties (and some host galaxy properties) show the interesting nature of very hard X-ray selected AGN. In order to understand the properties further, we are continuing to collect and analyze the properties in the optical through spectra (Winter et al. in prep) and imaging (Koss et al. in prep), the IR through {\\it Spitzer} observations (Weaver et al. in prep), and radio (Sambruna et al. in prep). Additionally, BAT is continuing to discover more sources at fainter fluxes, with sensitivity increasing as $\\sqrt t$. With the additional sources in future BAT catalogs, we will obtain an even better understanding of local AGN." }, "0808/0808.1927_arXiv.txt": { "abstract": "We discuss the dynamics of microquasar jets in the interstellar medium, with specific focus on the effects of the X-ray binaries' space velocity with respect to the local Galactic standard of rest. We argue that, during late stages in the evolution of large scale radio nebulae around microquasars, the ram pressure of the interstellar medium due to the microquasar's space velocity becomes important and that microquasars with high velocities form the Galactic equivalent of extragalactic head--tail sources, i.e., that they leave behind trails of stripped radio plasma. Because of their higher space velocities, low--mass X-ray binaries are more likely to leave trails than high--mass X--ray binaries. We show that the volume of radio plasma released by microquasars over the history of the Galaxy is comparable to the disk volume and argue that a fraction of a few percent of the radio plasma left behind by the X-ray binary is likely mixed with the neutral phases of the ISM before the plasma is removed from the disk by buoyancy. Because the formation of microquasars is an unavoidable by-product of star formation, and because they can travel far from their birth places, their activity likely has important consequences for the evolution of magnetic fields in forming galaxies. We show that radio emission from the plasma inside the trail should be detectable at low frequencies. We suggest that LMXBs with high detected proper motions like XTE J1118+480 will be the best candidates for such a search. ", "introduction": "\\label{sec:introduction} The interaction of AGN jets with their environments has been under investigation for several decades, partly because it is easily observable through the morphology of radio lobes \\citep[e.g.][]{miley:80} and X-ray cavities \\citep[e.g.][]{mcnamara:07}. Because the most powerful AGNs are essentially stationary in the centers of galaxy clusters, models of this interaction typically only consider stationary atmospheres that jets run into (see, e.g., \\citealt{heinz:06b} for recent work on jets in dynamic atmospheres). The evolution in this case can be separated into three distinct phases: (1) the early momentum driven phase, where the ram pressure of the jet is significant for the dynamical evolution and the source evolves into an elongated structure with narrow cocoons, (2) the energy driven phase, where the slowed-down jet plasma inflates supersonically expanding lobes, excavating a quasi-spherical cavity (this phase is well described by the self-similar solution by \\citealt{castor:75,kaiser:97,heinz:98}), and (3) the late, sub-sonic evolution, when the radio lobes (the reservoirs of relativistic gas released by jets) are in pressure equilibrium with the environment. In the case of AGNs, this radio plasma is buoyant in the host galaxy/galaxy cluster atmosphere \\citep[see][for a more detailed review]{reynolds:02}. It is becoming increasingly clear that X-ray binaries (XRBs) are also producing relativistic jets over a wide range in accretion rate \\citep{fender:00,fender:01,gallo:03,fender:04}. The inner regions of these XRB jets, where their dynamics is governed by the atmosphere of the compact object powering them, are very similar to the inner regions of AGN jets \\citep{heinz:03a}. However, \\cite{heinz:02c} showed that some critical differences exist between XRB jets and AGN jets in how they interact with the larger scale environment, well outside of the sphere of influence of the central compact object. One of the differences is that, in comparison with AGN jets, the ISM poses a much weaker barrier to microquasar jets. This is because the XRB jet thrust (in other words, the ram pressure delivered by the jet) is much larger in comparison to the inertial density of the ISM than it is in the case of AGN jets. The second fundamental difference between the two cases is, of course, that XRBs do not reside in the centers of dark matter halos with stratified gaseous atmospheres. Instead, they travel through regular Galactic ISM with some space velocity $v_{\\rm XRB}$, set by supernova kicks and orbital dynamics. The velocity dispersion of XRBs implies that a significant fraction of these sources are moving with large (supersonic) velocities through the ISM, which will have important consequences for their dynamics. VLBI parallax measurements have shown that a microquasars can be moving with velocities in excess of $100\\,{\\rm km\\,s^{-1}}$ with respect to the local standard of rest \\citep{mirabel:01}. As a result, the ram pressure of the ISM will act on the radio plasma released by the source, sweeping back the outer layers of the radio lobe. In this paper, we argue that this aspect has important consequences for the interaction of some XRB jets with their environment: Instead of inflating stationary cocoons, in many cases they will produce trails of radio emitting plasma (made up of relativistic particles and magnetic fields) as they travel through the Galaxy. A very similar situation is encountered in pulsar bow shock nebulae \\citep{cordes:96,frail:96}, where a relativistic wind inflates a channel in the ISM through which the pulsar is moving, and, of course, in exragalactic head-tail sources, which are the direct equivalent of jets propagating into a moving medium. The paper is organized as follows: In \\S\\ref{sec:dynamics}, we will present a simple parametric model of the interaction of the jets with the ISM. In \\S\\ref{sec:discussion} we discuss the consequences for particle and magnetic field input into the ISM, \\S\\ref{sec:observations} discussed the observational signatures of this interaction, and \\S\\ref{sec:summary} presents a brief summary of the paper. ", "conclusions": "\\label{sec:summary} We proposed that jets from X-ray binaries (microquasars) leave behind trails of non-thermal, synchrotron emitting plasma as they move through the interstellar medium. LMXBs are more likely to leave such trails due to their longer life expectancy and, more importantly, due to the higher expected space velocities, while HMXBs likely produce more stationary radio lobes. The total plasma volume deposited by XRBs over the history of the Galaxy is comparable to the total disk volume and constitutes a non-negligible fraction of the halo volume. We argue that a fraction $f_{\\rm mix}$ of order a few percent of this plasma is mixed into the thermal ISM. The magnetic field thus released can easily provide the seed field for Galactic dynamos to produce the fields observed in spiral galaxies even under the most conservative assumptions. Since HMXBs turn on rapidly after an episode of star formation, this mechanism should be important for galactic magnetic field production and maintenance within a time frame of about $10^{7}\\,{\\rm yrs}$ from the first supernovae. We showed that radio emission from the plasma inside the trail should be visible primarily at low frequencies, since the radiative cooling time of the plasma at GHz frequencies limits emission to a roughly spherical region around the binary. LOFAR, MWA, LWA and other future low frequency, wide field instruments are ideally suited for searches of this emission. LMXBs with high detected proper motions like XTE J1118+480 will be the best candidates for such a search." }, "0808/0808.1050_arXiv.txt": { "abstract": "Various laboratory-based experiments are underway attempting to detect dark matter directly. The event rates and detailed signals expected in these experiments depend on the dark matter phase space distribution on sub-milliparsec scales. These scales are many orders of magnitude smaller than those that can be resolved by conventional N-body simulations, so one cannot hope to use such tools to investigate the effect of mergers in the history of the Milky Way on the detailed phase-space structure probed by the current experiments. In this paper we present an alternative approach to investigating the results of such mergers, by studying a simplified model for a merger of a sub-halo with a larger parent halo. With an appropriate choice of parent halo potential, the evolution of material from the sub-halo can be expressed analytically in action-angle variables, so it is possible to obtain its entire orbit history very rapidly without numerical integration. Furthermore by evolving backwards in time, we can obtain arbitrarily-high spatial resolution for the current velocity distribution at a fixed point. Although this model cannot provide a detailed quantitative comparison with the Milky Way, its properties are sufficiently generic that it offers qualitative insight into the expected structure arising from a merger at a resolution that cannot be approached with full numerical simulations. Preliminary results indicate that the velocity-space distribution of dark matter particles remains characterized by discrete and well-defined peaks over an extended period of time, both for single and multi-merging systems, in contrast to the simple smooth velocity distributions sometimes assumed in predicting laboratory experiment detection rates. In principle, this structure contains a wealth of information about the formation history of the Milky Way's dark halo. ", "introduction": "Dark matter (hereafter DM) appears to be the dominant mass component of galaxies and large-scale structures in the Universe. The first evidence came in the 1930s (Zwicky 1933, Smith 1936), but it was only in the 1970s that observations of the rotation curves of galaxies demonstrated that DM dominates the masses of galaxies (Rubin \\& Ford 1970, Rubin, Thonnard \\& Ford Jr 1980). These observations showed that many rotation curves are approximately flat, or even rising, in the outer region of galaxies, where there is little luminous matter and so a Keplerian decline is expected. Subsequently, work on the hierarchical structure formation paradigm showed that non-baryonic material known as ``cold dark matter'' (CDM) is required to match the observed large-scale structure of the Universe (Peebles 1982). The term ``cold'' derives from the fact that this material was non-relativistic at the epoch of matter-radiation equality. The density of this CDM has subsequently been indirectly measured by various experiments such as 2dFGRS (Percival et al., 2001), WMAP (Dunkley et al. 2008) and the Sloan Digital Sky Survey (Tegmark et al. 2006). Particle physics provides us with various well-motivated candidates for the CDM, including weakly interacting massive particles (WIMPs). WIMPs can potentially be directly detected in the laboratory via their elastic scattering on target nuclei, and numerous experiments are currently underway to try to detect this phenomenon (e.g. Angle et al. 2007, Ahmed et al. 2008). The signals expected in these experiments, the number of recoil events per unit energy (and in some case its temporal and angular dependence), depend on the WIMP velocity distribution in the solar neighbourhood. A single stream of dark matter particles produces a step in the energy spectrum, detectable by a detector. The energy at which the step occurs is determined by the speed of the particles composing the stream (in the rest frame of the detector), while the height and position of the step vary annually, due to the Earth's orbit. Multiple streams would lead to a more complicated picture, and a superposition of enough such streams would ultimately be indistinguishable from a smooth distribution, depending on the detector resolution. The question that we seek to address here is what form one might expect for this distribution in reality. Predictions of the expected signals are often based on simplified models, which assume, for example, that the WIMP velocity distribution is Maxwellian (Freese, Frieman \\& Gould 1988) or a multivariate Gaussian (Evans, Carollo \\& de Zeeuw 2000, Helmi, White \\& Springel 2002). These models rely on the assumption that the Milky Way halo has reached a steady state so that the ultra-local DM phase-space distribution is smooth. However since structures form hierarchically, and the age of the Universe is not large compared with relevant dynamical timescales (such as the crossing time), this assumption is somewhat questionable. Numerous N-body simulations have been performed studying the hierarchical formation and evolution of DM halos. Such simulations find that DM halos contain ubiquitous substructure, in particular in their outer regions (Moore et al.\\ 1999, Klypin et al.\\ 1999). However, the salient question here is whether or not the DM distribution is smooth on the scales probed by direct detection experiments. Unfortunately, the resolution of even the best N-body simulations is many orders of magnitude larger than the relevant scales. The Sun's circular velocity around the centre of the Galaxy is $v_{\\odot} \\approx 200 \\; \\rm{km/sec}$, so that over a course of a year a terrestrial DM detector travels a distance \\begin{equation} \\rm{r_{det}} \\approx v_{\\odot} \\; \\tau_{\\rm{exp}} \\sim (200 \\; \\rm{km/sec}) \\; ({1 \\; \\rm{yr}}) \\sim 0.1 \\; \\rm{mpc} \\,, \\end{equation} whereas current N-body simulations cannot resolve scales smaller than ${\\sim \\; 100 \\, \\rm pc}$ (e.g. Diemand et al. 2007). This represents an insurmountable problem for the conventional simulation techniques, indicating that a completely different, specialized approach is necessary to describe in detail the ultra-fine DM distribution probed by direct detection experiments. A first attempt at such a specialized simulation was carried out by Stiff and Widrow (2003; hereafter SW). Their method used a reverse simulation process to calculate the DM speed distribution, $f(v)$, at a single spatial point of the phase space, representing a detector. More specifically they ran a simulation of the formation of a DM halo and at the end of the simulation put down a uniform grid (in velocity space) of mass-less test particles at the point of interest. They then evolved the test and simulation particles back to the initial time, found where the phase-space sheet of the test particles intersects the initial DM phase-space distribution, and hence calculated the density of the test particles at the final detector position. In this way, they found that the DM distribution in the solar neighbourhood is characterized by a number of discrete peaks. This numerical approach allowed them to use initial conditions which reproduce a realistic hierarchical-formation model for the Milky Way. However, a drawback of this technique was that it proved numerically unstable, and, in order to stabilize the reverse integration, SW were obliged to introduce a softening length of some $20\\, {\\rm kpc}$ into the gravitational force law that they applied. This softening is worryingly large as it significantly exceeds the solar radius in the Milky Way, $R_0 \\approx 8.5\\, {\\rm kpc}$, so might be expected to affect the inferred DM phase space distribution impinging on a terrestrial detector. More recently Vogelsberger et al. (2008) have formulated a technique for calculating the evolution of the fine-grained DM density in both static potentials and N-body simulations. They argue that the small-scale DM distribution that a terrestrial experiment would observe can be described by a multivariate Gaussian, in apparent contradiction to the SW results, perhaps indicating that the SW analysis had been compromised by the modification that they had to make to the gravitational force. In this paper we develop a complementary approach to the analysis of the ultra-fine DM distribution. Following SW, we calculate the DM distribution in the solar neighbourhood via a backward evolution method, but using a simplified model for the potential that allows the system to be expressed in action-angle (hereafter AA) variables. Although this simplified potential provides a less realistic representation of the Milky Way, its qualitative properties are similar, and it has the great benefit of being analytically soluble. Thus, the gravitational force does not have to be artificially softened (allowing one to test whether this effect did compromise the SW results), and one can very rapidly explore parameter space without the computational overhead of numerical integration. The paper is organized as follows. In Section~2 we present the simplified model that we use to describe the interaction between a galaxy like the Milky Way and a merging sub-halo. Section~3 contains our initial results, and we conclude in Section~4 with a discussion. ", "conclusions": "" }, "0808/0808.1099_arXiv.txt": { "abstract": "We report the results of a search for pure rotational molecular hydrogen emission from the circumstellar environments of young stellar objects with disks using the Texas Echelon Cross Echelle Spectrograph (TEXES) on the NASA Infrared Telescope Facility and the Gemini North Observatory. We searched for mid-infrared H$_{2}$ emission in the \\sone, \\stwo, and \\sfour ~transitions. Keck/NIRSPEC observations of the H$_{2}$ \\snine ~transition were included for some sources as an additional constraint on the gas temperature. We detected H$_{2}$ emission from 6 of 29 sources observed: AB Aur, DoAr 21, Elias 29, GSS 30 IRS 1, GV Tau N, and HL Tau. Four of the six targets with detected emission are class I sources that show evidence for surrounding material in an envelope in addition to a circumstellar disk. In these cases, we show that accretion shock heating is a plausible excitation mechanism. The detected emission lines are narrow ($\\sim$10 km s$^{-1}$), centered at the stellar velocity, and spatially unresolved at scales of 0.4\\arcsec, which is consistent with origin from a disk at radii 10-50 AU from the star. In cases where we detect multiple emission lines, we derive temperatures $\\gtrsim$ 500 K from $\\sim$1 M$_{\\oplus}$ of gas. Our upper limits for the non-detections place upper limits on the amount of H$_2$ gas with T $>$500~K of less than a few Earth masses. Such warm gas temperatures are significantly higher than the equilibrium dust temperatures at these radii, suggesting that the gas is decoupled from the dust in the regions we are studying and that processes such as UV, X-ray, and accretion heating may be important. ", "introduction": "Studying the structure and evolution of circumstellar disks is crucial to developing an understanding of the process of planet formation. Observations of dust emission and modeling of the spectral energy distributions (SED) of disks have revealed much about the dust component from a few stellar radii out to hundreds of AU \\citep{zuckerman01}. While circumstellar disks are composed of both dust and gas, the gas component dominates the mass of the disk, with molecular hydrogen (H$_{2}$) being the most abundant constituent. In order to develop a complete picture of the structure and evolution of protoplanetary disks, it is important to observe the gas. Observations at different wavelengths probe different disk radii. Submillimeter observations sample gas at large radii ($> 50$ AU) \\citep{semenov05}, while near-infrared CO \\citep{najita03,blake04} and H$_{2}$O \\citep{carr04, thi05} observations allow for study of the inner few AU. Spectral lines in the mid-infrared (5-25 $\\mu$m) provide a means to investigate gas in the giant planet region of the disk and beyond (10-50 AU) \\citep{najita07a}. Several mid-infrared spectral diagnostics have been shown to be useful probes of gas in disks. These include [NeII] at 12.8 $\\mu$m \\citep{pascucci07,lahuis07,herczeg07}, H$_{2}$O rotational transitions \\citep{carr08,salyk08}, [FeI] at 24 $\\mu$m \\citep{lahuis07}, and, based on a theoretical analysis of debris disks, [SI] at 25.2 $\\mu$m, and [FeII] at 26 $\\mu$m \\citep{gorti04}. Molecular hydrogen should make up the bulk of the mass in disks, but is a challenge to detect. Bright far-ultraviolet (FUV) H$_{2}$ emission from classical T Tauri stars may be produced in the irradiated disk surface \\citep{herczeg02,bergin04}. At longer wavelengths, rovibrational and pure rotational transitions are generally weak because H$_2$ lacks a permanent dipole moment. Near-infrared emission in the \\textit{v}=1-0 \\sone ~rovibrational transition of H$_{2}$ has been detected from T Tauri stars \\citep{bary03,ramsay07,carmona08b} and may be the result of excitation by UV and X-ray irradiation \\citep{nomura07,gorti08}. Near-infrared adaptive optics fed, integral field spectroscopy of six classical T Tauri stars that drive powerful outflows has revealed that most of the H$_{2}$ emission is spatially extended from the continuum \\citep{beck08}. The properties of the emission are consistent with shock excitation from outflows or winds rather than UV or X-ray excitation from the central star. The FUV H$_{2}$ emission probes gas between 2000 and 3000 K \\citep{herczeg04,nomura05} and the 2.12 $\\mu$m H$_{2}$ line traces gas at T$>$1000 K \\citep{bary03}. The mid-infrared H$_{2}$ lines considered in this paper are most sensitive to gas at lower temperatures. Owing to their small Einstein A-values, the pure rotational mid-infrared H$_{2}$ lines remain optically thin to large column densities (N$_{H_{2}} > 10^{23}$ cm$^{-2}$) and will be in LTE at the densities found in disks. For a disk with a strong mid-infrared continuum, the dust becomes optically thick well before the H$_{2}$ lines. Observable line emission is present only when there is a temperature inversion in the atmosphere of the disk or if there is a layer of dust-depleted gas separate from the optically thick dust. The process of dust coagulating into larger grains or settling out of the disk atmosphere can allow a larger column of gas to be observed. A disk that has an optically thin mid-infrared continuum, implying a very small amount of dust in the disk or dust grains that have grown large compared to mid-infrared wavelengths, would allow the entire disk to be observable. However, it is not known whether such disks have large quantities of gas and, in the absence of gas heating through collisions with dust grains, another heating mechanism is necessary in dust-free environments, such as UV or X-ray heating \\citep{glassgold04,gorti04,nomura07}. A number of groups have searched for H$_{2}$ emission from protoplanetary disks in recent years. \\citet{thi01} reported the detection of several Jupiter masses of warm gas in a sample of disk sources based on \\textit{Infrared Space Observatory} (ISO) observations of the H$_{2}$ \\szero ~and \\sone ~lines. However, follow-up observations from the ground with improved spatial resolution did not confirm these results \\citep{richter02,sheret03,sako05}. \\citet{richter02} used the Texas Echelon Cross Echelle Spectrograph (TEXES) on the NASA 3m Infrared Telescope Facility (IRTF) to set upper limits on the warm gas mass within the disks of six young stars. \\citet{sheret03} searched for H$_{2}$ emission using MICHELLE on the United Kingdom Infrared Telescope and set upper limits on the emission from disks around two stars. The first group to use an 8-meter class telescope in the search for molecular hydrogen was \\citet{sako05} using the Cooled Mid-Infrared Camera and Spectrometer on the 8.2 m Subaru telescope to set upper limits for emission in the \\sone ~line around four young stars. Using the Infrared Spectrograph (IRS) aboard the \\textit{Spitzer Space Telescope}, \\citet{pascucci06} reported the non-detection of H$_{2}$ lines in their sample of 15 young Sun-like stars, while \\citet{lahuis07} detected the \\stwo ~and \\sthree ~lines in $\\sim$8\\% of the 76 circumstellar disks in their sample. Recently, ground-based observations with high resolution spectrometers on 8-meter class telescopes have produced both detections of the mid-infrared H$_{2}$ lines in the Herbig Ae stars AB Aur and HD 97048 \\citep{bitner07,martin07} and stringent upper limits in a sample of six Herbig Ae/Be stars and one T Tauri star \\citep{carmona08a}. Three mid-infrared pure rotational H$_{2}$ lines are observable from the ground: \\sone ~($\\lambda = 17.035 ~\\mu$m), \\stwo ~($\\lambda = 12.279 ~\\mu$m), and \\sfour ~($\\lambda = 8.025 ~\\mu$m). When multiple optically thin lines are observed, line ratios permit the determination of the temperature and mass of the emitting gas. Ratios of these three lines are most sensitive to temperatures of 200-800 K. Two additional pure rotational H$_{2}$ lines are observable near 5 $\\mu$m: \\seight ~($\\lambda = 5.053 ~\\mu$m) and \\snine ~($\\lambda = 4.695 ~\\mu$m), extending our temperature sensitivity to hotter gas. The high spectral resolution possible with an instrument like TEXES (Lacy et al. 2002) increases our sensitivity to narrow line emission and helps determine the location of the emission. By making observations at high spectral resolution, we maximize the line to continuum contrast while minimizing atmospheric effects by separating the lines from nearby telluric features. A further benefit of high spectral resolution is that we are able to estimate the location of the emitting gas if coming from a disk under the assumption of Keplerian rotation. In this paper, we present the results of a search for molecular hydrogen emission in disk sources using TEXES on both the IRTF and Gemini North telescopes. We observed 29 sources spanning a range of mass, age, and accretion rate in order to constrain the amount of warm gas in the circumstellar disks of these stars. ", "conclusions": "We have carried out a survey for pure rotational H$_{2}$ emission from the circumstellar environments surrounding a sample of 29 stars with disks and detected emission from 6. In the case of non-detections, our upper limits constrain the amount of T $>$ 500 K gas in the surface layers of the circumstellar disks to be less than a few Earth masses. Several objects in our survey have transition object SEDs implying the presence of an optically thin, dust-depleted inner disk: DoAr 21, GM Aur, HD 141569, and LkCa 15. Among these sources, only DoAr 21 shows H$_{2}$ emission and it appears to be far from the star. One possible explanation for transitional SEDs is grain growth \\citep{strom89,dullemond05} whereby the inner disk becomes optically thin yet remains gas rich. This gas could be heated through accretion as well as X-ray and UV heating. GM Aur is of particular interest since it has an accretion rate ($\\sim$10$^{-8}$ M$_\\odot$ yr$^{-1}$) similar to typical CTTS accretion rates so should have as much gas in the inner disk as a typical CTTS. If grain growth is responsible for the transitional SEDs of these sources, the available heating mechanisms are insufficient to produce detectable H$_{2}$ line emission. Alternatively, grain growth may not be a good explanation for a transition object SED, as suggested by other demographic data \\citep{najita07b}. In all cases, the detected emission lines are narrow and centered at the stellar velocity. The narrow range of line widths, FWHM between 7 and 15 km s$^{-1}$, along with the fact that the line fluxes are all similar, suggests that the mechanism for exciting the emission may be the same in each case. Four of the six targets with detected emission are class I sources that show evidence for surrounding material in an envelope in addition to a circumstellar disk. It is possible, and likely in the case of HL Tau, that the H$_{2}$ emission we observe is a result of gas in the circumstellar envelope being shock heated by an outflow. However, the fact that all of the H$_{2}$ line centroids in our sample are within a few km s$^{-1}$ of the stellar velocity argues against this being the case for all of our detections. Under the assumption of emission from a disk in Keplerian rotation, the narrow line widths imply that the emission arises at disk radii from 10-50 AU. At such large disk radii, additional heating of the gas besides heating due to collisions with dust grains is required to explain the temperatures derived from our H$_{2}$ observations. Both X-ray/UV irradiation of the disk surface layer and accretion shocks resulting from matter infall onto the disk are plausible candidates. With the exception of DoAr 21, all of the sources where we detect H$_{2}$ emission possess both a circumstellar disk and a surrounding envelope of material. This lends support to the possibility that the H$_{2}$ emission we observed may be the result of shocks in the disk due to infalling material. Models of molecular hydrogen emission from disks that assume sufficient levels of stellar X-ray and UV irradiation \\citep{gorti08} predict line fluxes that are consistent with our observations. In contrast, models which assume smaller values of stellar UV and X-ray irradiation \\citep{nomura07} produce weaker H$_{2}$ emission than observed in our sample. We looked for evidence of a correlation between X-ray/UV luminosity and the presence of H$_{2}$ emission but found none. We note that the X-ray and UV luminosities used for the purpose of searching for a correlation with H$_{2}$ emission were not measured at the same time. To definitively test for a correlation between X-ray/UV luminosity and the presence of H$_{2}$ emission will require a series of coordinated observations." }, "0808/0808.1785_arXiv.txt": { "abstract": "In this paper, using 2MASS photometry, we study the structural and dynamical properties of four young star clusters viz. King~16, NGC~1931, NGC~637 and NGC~189. For the clusters King~16, NGC~1931, NGC~637 and NGC~189, we obtain the limiting radii of $7'$, $12'$, $6'$ and $5'$ which correspond to linear radii of 3.6~pc, 8.85~pc, 3.96~pc and 2.8~pc respectively. The reddening values $E(B-V)$ obtained for the clusters are 0.85, 0.65--0.85, 0.6 and 0.53 and their true distances are 1786~pc, 3062~pc, 2270~pc and 912~pc respectively. Ages of the clusters are 6~Myr, 4~Myr, 4~Myr and 10~Myr respectively. We compare their structures, luminosity functions and mass functions ($\\phi(M) = dN/dM \\propto M^{-(1+\\chi)}$) to the parameter $\\tau = t_{age}/t_{relax}$ to study the star formation process and the dynamical evolution of these clusters. We find that, for our sample, mass seggregation is observed in clusters or their cores only when the ages of the clusters are comparable to their relaxation times ($\\tau \\geq 1$). These results suggest mass seggregation due to dynamical effects. The values of $\\chi$, which characterise the overall mass functions for the clusters are 0.96 $\\pm$ 0.11, 1.16 $\\pm$ 0.18, 0.55 $\\pm$ 0.14 and 0.66 $\\pm$ 0.31 respectively. The change in $\\chi$ as a function of radius is a good indicator of the dynamical state of clusters. \\bf{star clusters: young -- near-infrared photometry -- color--magnitude diagrams -- pre-mainsequence stars -- initial mass function--relaxation time-- 2MASS} ", "introduction": "Star clusters are the most fundamental stellar sytems in understanding the star formation process, which is still full of mysteries. A good study of these stellar systems is the basis of understanding galaxies which are the larger building blocks of the universe \\citep{lynga82,janes81,friel95,bonatto05,piskunov06}. \\end{abstract} Homogeneous samples of photometric data, coupled with uniform methods of data analysis are essential to make statistical inferences based on the fundamental parameters of clusters. These studies can contribute to understanding the galactic disk, formation and evolution of clusters, molecular cloud fragmentation, star formation and evolution. In this work, we study a sample of young clusters viz. King~16, NGC~1931, NGC~637 and NGC~189 using photometric data from the Two~Micron~All~Sky~Survey (2MASS) \\citep{Skrutskie06}. The 2MASS covers 99.99\\% of the sky in the near-infrared $J$~(1.25~$\\mu$m), $H$~(1.65~$\\mu$m) and $K_{s}$~(2.16~$\\mu$m) bands (henceforth $K_{s}$ shall be refered to as $K$). Hence the 2MASS database has the advantages of being homogeneous, all sky (enabling the study of the outer regions of clusters where the low mass stars dominate) and covering near infrared wavelengths where young clusters can be well observed in their dusty environments. \\cite{dutra01} discovered 42 objects at infra-red wavelengths using the 2MASS survey. Many papers devoted to the study of clusters using the 2MASS have been presented in the past few years \\citep{bica03,bica06,tadross07,kim06} showing the potential of this database. We use the 2MASS database to study the sample of four young clusters, to investigate the structure and dynamical state of these clusters close to their time of formation. The sample is selected on the basis that all the four clusters have an age of $\\approx$10~Myr reported in literature (see Table 1) and have been formed in different environments. We study the structures and dynamical states of our sample of clusters and determine their mass functions (MFs) and degree of mass seggregation in various regions of the clusters. To study the sample, we construct radial density profiles (RDPs), color--magnitude diagrams (CMDs), color--color diagrams, luminosity functions (LFs) and MFs. Such studies are not possible using heterogenous datasets where unknown biases may be present. The initial mass function (IMF) is the distribution of stars of varying masses from the original parent cloud. The universality of the IMF and the influence of environment on star formation is still a matter of debate. As these clusters are very young (age $\\leq$ 10 Myr), their MF may be approximated as the IMF. The sample of clusters have been made from differing initial conditions and subject to varied influences of external interactions with the galactic field, thus leading to observable differences, which we explore. However, from a recent study of \\cite{kroupa07}, even in the case of very young clusters, there is a change in the MF due to the dynamics of young clusters which loose a significant fraction of their stars at an early age. Mass segregation is the redistribution of stars according to their masses, thus leading to the concentration of high mass stars near the centre and the low mass ones away from the centre. This has been observed in a variety of clusters, both young and old. The variation of the MF of these clusters is determined in different regions of the clusters and their values are compared. Further, we estimate from the value of $\\tau = t_{age}/t_{relax}$, the degree of mass segregation expected due to dynamical effects and compare it with our observations. The relaxation time $t_{relax}$ is a characteristic time in which there is an equipartition of energy and the high mass stars with lesser kinetic energy sink to the core and the low mass stars move to the outer regions of the cluster \\citep{binney08}. The value of $\\tau$ indicates whether an excess of high mass stars in the cores of clusters is a result of dynamical evolution or the imprint of the star formation process itself. The parameter $\\tau$ has been described as an evolutionary parameter \\citep{bonatto05} which indicates the extent to which the cluster has relaxed. It relates to the core and overall MF flattening. For large values of $\\tau$, the high mass stars sink to the centre and the low mass stars with high velocities move towards the outskirts and hence the MFs of clusters show large-scale mass segregation and low-mass stars evaporation. We report the presence of gaps in the main sequence associated to physical processes in stars \\citep{kjeldsen91}. The plan of the paper is as follows: Section 2 describes the clusters in our sample and shows the corresponding RDPs and the values obtained for the limiting radii for these clusters. Section 3 describes the method of selecting cluster members and the corresponding values of fundamental parameters obtained. LFs and MFs are described in Section 4 and a comparative study of these clusters is in the concluding Section 5. ", "conclusions": "In this paper we have studied four young clusters of comparable ages to understand their structure and dynamics. The RDPs of the clusters have been plotted and the parameters for the clusters such as reddening, distance and age have been determined using isochrone fits. We have also plotted the LFs in the $J$, $H$ and $K$ bands and used the derived mass--luminosity relation to find the MFs using all three bands independently. The $\\chi$ values have been determined for different regions and the overall clusters as a function of the parameter $\\tau$. We use the difference in $\\chi$ values to estimate the level of mass segregation of the clusters and their cores. In the case of King~16, where $\\tau$ = 3 for the core, the core is clearly relaxed as in indicated by its flat $\\chi=-0.44$, while for the outer regions where $\\tau$ = 0.57, the cluster has begun to relax. In the case of NGC~1931, the core is relaxed and the outer region seems to have begun relaxation. In the case of NGC~637, the core is relaxed ($\\tau=4/0.039$), and there appears to be a redistribution of stars in the cluster, indicated by a progressive increase in $\\chi$ from 0.39 to 1.15. The outer halo has a larger value of $\\chi$ compared to the overall cluster (0.55) as, it appears that the inner halo has thrown away a large number of low mass stars to the outer halo, thus causing an excess in low mass stars and a steeper $\\chi$. In the case of NGC~189, where $\\tau$ =2.79, the cluster has already relaxed, as is indicative of the flat $\\chi$ of the core. The outer core has a flatter $\\chi$, probably because low mass stars which were thrown out of the inner halo during relaxation, have been lost and hence the outer halo has a deficit of low mass stars. As, seen from our analysis, mass seggregation is observable in the cores of the clusters King~16, NGC~1931 and NGC~637 where the cluster ages are comparable to the relaxation times. In the case of NGC 189, where the relaxation time is lesser than the age of the cluster, we notice a flatter mass function. This implies that the observed mass seggregation in these clusters is a dynamical effect. However, larger samples will improve the statistics and give us a better insight in the physical processes leading to the structure and dynamical evolution of clusters." }, "0808/0808.2053_arXiv.txt": { "abstract": "Infrared Dark Clouds (IRDCs) represent the earliest observed stages of clustered star formation, characterized by large column densities of cold and dense molecular material observed in silhouette against a bright background of mid-IR emission. Up to now, IRDCs were predominantly known toward the inner Galaxy where background infrared emission levels are high. We present {\\it Spitzer} observations with the Infrared Camera Array toward object G111.80+0.58 (G111) in the outer Galactic Plane, located at a distance of $\\sim$3\\,kpc from us and $\\sim$10\\,kpc from the Galactic center. Earlier results show that G111 is a massive, cold molecular clump very similar to IRDCs. The mid-IR {\\it Spitzer} observations unambiguously detect object G111 in absorption. We have identified for the first time an IRDC in the outer Galaxy, which confirms the suggestion that cluster-forming clumps % are present throughout the Galactic Plane. However, against a low mid-IR background such as the outer Galaxy it takes some effort to find them. ", "introduction": "Massive stars are believed to form almost exclusively in stellar clusters \\citep{Blaauw1964,deWit2005}, where supposedly the majority of all stars in the Galaxy form. The precursors to these stellar clusters are cold, dense and, most importantly, massive molecular clumps. A major, unexpected and exciting result from the {\\it Midcourse Space Experiment} ({\\it MSX}) and {\\it Infrared Space Observatory} ({\\it ISO}) is the discovery of dark clouds seen in silhouette against the bright mid-IR background toward the inner Galactic Plane \\citep{MSXIRDC1998,ISOCAM1996}. In contrast to the well-known dark clouds seen in visual extinction against the stellar distribution, these so-called Infrared Dark Clouds (IRDCs) have much higher column densities ($\\gtrsim$\\,10$^{23}$\\,cm$^{-2}$), are % at large distances ($\\gtrsim$\\,1\\,kpc) and hence are typically much more massive. Follow-up studies \\citep[e.g.][]{Carey1998,Teyssier2002,Simon2006co} show that they are cold ($<$\\,20\\,K), dense ($\\sim$\\,10$^5$\\,cm$^{-3}$) and, indeed, among the most massive molecular clumps yet found in our Galaxy ($M$$\\sim$100--10$^5$\\,M$_{\\odot}$). Many IRDCs contain compact (sub-) millimeter cores \\citep[$M$$\\sim$10--10$^3$\\,\\msun, $n$$\\sim$10$^3$--10$^7$\\,cm$^{-3}$, $T$$\\sim$15--30\\,K, e.g.,][]{Carey2000,Teyssier2002,Garay2004,Rathborne2006}. While these cores were first considered to represent the pre-stellar core phase of clustered star formation, recent studies show that at least some cores in IRDCs contain % proto-stellar objects \\citep[e.g.,][]{Redman2003,Ormel2005,Rathborne2005}. Nevertheless, IRDCs are likely to represent the earliest stages of clustered star formation \\citep[e.g.,][]{Menten2005,Rathborne2006} and they are generally referred to as cluster-forming clumps. Their ambient physical conditions can illuminate the role of IRDCs in the star forming process and can provide insight into the differences between low- and high-mass star formation, the nature of the initial mass function and the impact of environment on star formation. Moreover, their % core forming properties may reveal the important mechanisms involved in forming high-mass stars, e.g., competitive accretion, merging of lower-mass stars or massive disk accretion \\citep[e.g.,][and references therein]{Larson2007}. The identification of IRDCs is by necessity biased toward cold molecular clouds having sufficient column density against a bright mid-IR background, in order to be seen as absorption features. Hence, identifying IRDCs is considerably easier toward the mid-IR bright spiral arms and molecular ring in the inner Galactic Plane \\citep{Simon2006msx}, i.e., within 90\\degree\\ from the Galactic Center. An unbiased, more complete understanding of clustered star formation and its dependence on environment and other external properties, e.g., interstellar radiation field, metallicity, external pressure and dynamics, requires a study of IRDC-like objects in more quiescent regions, such as the outer Galaxy, i.e., between $l$=90\\degree and $l$=270\\degree. Frieswijk et al. (2008) produce a catalog of extended red objects in the outer Galactic Plane. The objects are identified in the {\\it Two Micron All Sky Survey Point Source Catalog} ({\\it 2MASS,} \\citealt{Skrutskie2006}) as clusters of stars that show a statistically redder color distribution compared to their local surroundings. A sample, consisting of 414 objects covered by the Canadian Galactic Plane Survey \\citep[CGPS,][]{Heyer1998}, was investigated in detail using the CO data of the CGPS. Over 90\\% of the objects correlate morphologically with CO emission with a typical FWHM of the order of 5\\,km\\,s$^{-1}$. This suggests that the observed reddening is due to the presence of foreground molecular clouds. The kinematic distance of the clouds, derived from the CO radial velocity, indicates that $\\sim$15\\% are located beyond $\\sim$\\,3\\,kpc from us. Some of these distant clouds ($\\sim$10) do not show significant mid- and far-IR emission in, e.g., MSX and IRAS, and are potentially cold, cluster-forming clumps in the outer Galaxy, or `IRDC-like' objects. \\citet[][F07]{Frieswijk2007AA} studied the physical characteristics of a candidate IRDC-like object (G111) by using molecular line observations. At a radial velocity of -52\\,km\\,s$^{-1}$, G111 is located at a kinematic distance of $\\sim$\\,5\\,kpc, assuming IAU standard constants. However, within 15\\arcmin\\ on the sky and at similar radial velocity are the (massive) star forming regions NGC 7538 and S 159. These are part of the Cas OB2 complex in the Perseus spiral arm, located at $\\sim$3\\,kpc from us and $\\sim$10\\,kpc from the Galactic center. Hence, a distance of 3\\,kpc is assumed for G111. A detailed investigation of the molecular lines conducted by F07 revealed multiple cold and massive cores (P1--P10, see Fig.\\ref{fig1} and Sec.\\ref{conclusions}) along a filamentary cloud structure. The main conclusion is that G111 has global properties in agreement with values found for IRDCs, e.g., $M$\\,$\\gtrsim$\\,3000\\,\\msun, $N$\\,$\\gtrsim$\\,10$^{23}$\\,cm$^{-2}$, $T$\\,$\\lesssim$\\,20\\,K. The mass and density of individual cores are low when compared to sub-millimeter cores in IRDCs. However, with peak column densities in excess of 10$^{23}$\\,cm$^{-2}$, the most massive cores in G111 should, provided that sufficient mid-IR background is present, be observable as absorption features in a similar way to inner Galaxy IRDC cores. We subsequently requested {\\it Spitzer} observations toward G111 to test the idea that the cores can be observed in absorption. {\\it MSX} observations were not able to resolve this, presumably because the low sensitivity did not allow a detection of contrast between the molecular cloud and the background. With the sensitivity of the Infrared Camera Array (IRAC) on {\\it Spitzer} we argued that it should be possible to observe this contrast. In this paper, we present the first outcomes of the data. Combined with the results in F07, we conclude that we have, beyond doubt, a detection of an IRDC in the outer Galaxy. The paper is organized as follows: Section \\ref{observations} gives % an overview of the observations. Section \\ref{results} presents the results. In Section \\ref{conclusions} we discuss the observations and give our interpretation of the data % compared to the results published in F07. We conclude the section with some future prospects. ", "conclusions": "\\label{conclusions} The mid-IR properties of G111 compare well to other IRDCs. The {\\it MSX} survey data have been used by \\citet{Simon2006msx} to identify a large number of IRDCs in the inner Galaxy by their mid-IR contrast relative to the background. A typical 8\\,$\\mu$m extinction observed toward IRDCs, determined from the ratio between the on- and off-source emission after foreground and zodiacal light correction, is 1--2\\,mag \\citep[e.g.,][]{MSXIRDC1998,Carey2000}. We find extinction values toward G111 to be in agreement with this, at least toward the darkest cores. More so, the extinction measurement gives only a strict lower limit to the column of material. The cloud does not only absorb background emission, but re-radiates low-level emission as well. For a cold dark cloud ($<$25\\,K) this will be negligible. However, once embedded objects start heating parts of the cloud it may become important. The contrast between the cloud and the background, and thus the measured extinction, decreases as the cloud emits, but of course the column of material does not. The molecular column densities of positions P1--P10 are $\\sim$10$^{22}$\\,cm$^{-2}$, peaking at $\\approx$\\,7$\\times$10$^{22}$\\,cm$^{-2}$ for P8 (F07). The latter corresponds to 2.8\\,mag extinction at 8\\,$\\mu$m (70\\,mag in $A_V$)\\footnote{This is a factor of 2 above the value in F07, because of a too low conversion factor used by F07.}. These estimates demonstrate that parts of this cloud have a mid-IR extinction comparable to values found for inner Galaxy IRDCs \\citep[e.g.,][]{Carey1998,Hennebelle2001}. Recall in this that the selection of IRDCs from {\\it MSX} data is based on high contrast \\citep{Simon2006msx}. In the outer Galaxy, a high contrast is not guaranteed because of the low-level background emission. For example, compare the background of the IRDCs toward the edge of W51, i.e., $\\sim$\\,60\\,MJy\\,sr$^{-1}$ (IRAC/MIPS cycle 1 observations, K. Kraemer et al., priv. comm.) with the background that we observe ($\\sim$\\,20\\,MJy\\,sr$^{-1}$). Note that one bright 8\\,$\\mu$m emission feature clearly is in the background of G111, since a dark lane is visible in front of IRAS23136+6111 (see Figure \\ref{fig1}, lower blow-up). The four-color images show a clustering of star-like objects near the dark filaments. Though these stars could coincidently be in the foreground, this positional consistency suggests that some cores might have started to form stars. The excess in steller surface density was also pointed out by F07 toward P5 and P8, based on the 2MASS data. Note that IRDCs are often associated with active star formation \\citep[e.g.,][]{Redman2003,Rathborne2005}. Star forming activity may further reveal itself by the presence of `green fuzzy emission', i.e., as weak extended 4.5\\,$\\mu$m features. This emission is often attributed to shocked gas arising from outflow-activity, e.g., the pure rotational H$_2$ lines S(11) at 4.18\\,$\\mu$m through S(9) at 4.69\\,$\\mu$m and the ro-vibrational lines of CO at 4.45--4.95\\,$\\mu$m \\citep[e.g.,][]{Noriega2004,Marston2004}. Shocked gas features are, besides for nearby star formation, also frequently observed toward IRDCs \\citep[e.g.,][]{Rathborne2005,Beuther2007}. A first impression of the emission at 4.5\\,$\\mu$m suggests that such features are present toward some cores in G111. However, further investigation is required to confirm this. \\\\ \\ \\\\ \\textbf{Notes on individual cores}\\\\ The positions P1 to P10 in Figure \\ref{fig1} were selected by F07 based on their high column density. A detailed investigation showed that they represent cold (10--20\\,K), dense ($>$10$^3$\\,cm$^{-3}$) and massive cores ($\\sim$100\\,\\msun, P8$\\sim$\\,1000\\,\\msun) where stars might form. Star forming activity was investigated by means of 2MASS star-colors and -counts and the presence of warm gas (NH$_{3}$, $^{13}$CO). The following discussion summarizes some of the results in F07 and includes the mid-IR characteristics of the cores presented in this paper. P1, P2, P3 and P7 do not show indications of active star formation. P1 shows enhanced 8\\,$\\mu$m emission, which may be a chance encounter along the line of sight. The other cores show a decrease in 8\\,$\\mu$m emission, corresponding to an $A_V$ of $>$10\\,mag. The peak extinction appears somewhat offset from the C$^{18}$O peak, which is true for most cores. This may be explained by C$^{18}$O freeze-out in the densest regions. P4 is not a core but part of the C$^{18}$O filament extending to the east. Several extinction peaks at 8\\,$\\mu$m ($A_V$$\\gtrsim$10\\,mag) are present. P9 and P10 are not evident as cores in C$^{18}$O and show no specific features at 8\\,$\\mu$m. P6 shows a decrease of 8\\,$\\mu$m emission ($A_V$$\\gtrsim$7\\,mag), but the contrast with the background may be reduced due to the presence of the bright source near the C$^{18}$O peak. Whether this source is physically associated cannot be determined at present. P5 and P8 show signs of active star formation. The NH$_{3}$ lines indicate the presence of warm gas. Both cores have associated 2MASS sources with typical near-IR colors of YSOs. Furthermore, the stellar surface density of 2MASS peaks at a value almost twice that of the surrounding field for P8. These results are supported by the four-color images that reveal signs of active star formation in the form of clustering of objects and 4.5\\,$\\mu$m `green fuzzes'. The 8\\,$\\mu$m contrast indicates peak extinctions of at least 12.5\\,mag in $A_V$ for both cores. IRAS23136+6111 was not a target position in F07, but is nonetheless interesting to include here. C$^{18}$O emission reveals a narrow filamentary structure. The four-color image depicts a dark lane and a clustering of point sources matching the location of the filament. We thus conclude that it must be in the foreground of the bright IRAS object. Further investigation is required to see if the dark lane and the IRAS source are associated, which may indicate a triggered star forming event. To summarize, the {\\it Spitzer} observations support the results in F07. Except P5 and P8, where star formation may have started, the cores appear to be in a cold, pre-stellar phase. \\\\ \\ \\\\ \\textbf{Future prospects}\\\\ Supplemental IRAC and MIPS 24\\,$\\mu$m data from {\\it Spitzer} are expected. Combining the IRAC bands with the 2MASS data enables a discrimination between stars and YSOs for many of the clustered point sources in the cloud vicinity. Adding 24\\,$\\mu$m data will allow a proper SED modeling of the dust emission. The characteristic features of envelopes, disks and photospheres in the SED will enable a determination of the evolutionary state, i.e., Class 0, I, II or III, of the apparent discrete sources \\citep[e.g.,][]{Robitaille2007}. Star forming activity can be further characterized through an analysis of the spatial distribution and strength of polycyclic aromatic hydrocarbon emission (e.g., 3.3, 6.2 and 7.7\\,$\\mu$m covered by the 3.6, 5.8 and 8\\,$\\mu$m bands, respectively) and shock activity, exemplified by the 4.5\\,$\\mu$m features in the blow-up images in Figure \\ref{fig1} \\citep[e.g., bow shocks;][]{Neufeld2006,Velusamy2007}. Considering the distance to the cloud, high-resolution continuum and spectroscopic observations (near-IR to millimeter) are essential to improve on the spatial information." }, "0808/0808.2323_arXiv.txt": { "abstract": "\\noindent We discuss the 21cm power spectrum (PS) following the completion of reionization. In contrast to the reionization era, this PS is proportional to the PS of mass density fluctuations, with only a small modulation due to fluctuations in the ionization field on scales larger than the mean-free-path of ionizing photons. We derive the form of this modulation, and demonstrate that its effect on the 21cm PS will be smaller than 1\\% for physically plausible models of damped Ly$\\alpha$ systems. In contrast to the 21cm PS observed prior to reionization, in which HII regions dominate the ionization structure, the simplicity of the 21cm PS after reionization will enhance its utility as a cosmological probe by removing the need to separate the PS into physical and astrophysical components. As a demonstration, we consider the Alcock-Paczynski test and show that the next generation of low-frequency arrays could measure the angular distortion of the PS at the percent level for $z\\sim3-5$. ", "introduction": "Recently, there has been much interest in the feasibility of mapping the three-dimensional distribution of cosmic hydrogen through its resonant spin-flip transition at a rest-frame wavelength of 21cm~(Furlanetto, Oh \\& Briggs~2007; Barkana \\& Loeb~2007). Several experiments are currently being constructed (including MWA~\\footnote{http://www.haystack.mit.edu/ast/arrays/mwa/}, LOFAR~\\footnote{http://www.lofar.org/}, PAPER ~\\footnote{http://astro.berkeley.edu/~dbacker/EoR/}, 21CMA~\\footnote{http://web.phys.cmu.edu/~past/}, GMRT~\\footnote{Pen et al.~(2008)}) and more ambitious designs are being planned (SKA~\\footnote{http://www.skatelescope.org/}). One driver for mapping the 21cm emission is the possibility of measuring cosmological parameters from the shape of the underlying power spectrum (PS; see Loeb \\& Wyithe 2008). During the epoch of reionization, the PS of 21cm brightness fluctuations is shaped mainly by the topology of ionized regions, rather than by the PS of matter density fluctuations which is the quantity of cosmological interest~(McQuinn et al.~2006; Santos et al.~2007; Iliev et al.~2007). As a result, the line-of-sight anisotropy of the 21cm PS due to peculiar velocities must be used to separate measurements of the density PS from the unknown details of the astrophysics (Barkana \\& Loeb~2005; McQuinn et al.~2006). The situation is expected to be simpler both prior to the formation of the first galaxies (at redshifts $z\\ga 20$, Loeb \\& Zaldarriaga~2004; Lewis \\& Challinor~2007; Pritchard \\& Loeb~2008), and following reionization of the intergalactic medium (IGM; $1\\la z\\la6$) -- when only dense pockets of self-shielded hydrogen, such as damped Ly$\\alpha$ absorbers (DLA) and Lyman-limit systems (LLS) survive ~(Wyithe \\& Loeb~2008; Chang et al.~2007; Pritchard \\& Loeb~2008). In this paper we focus our discussion on the post-reionization epoch (Wyithe \\& Loeb~2007; Chang et al.~2007). The DLAs which contain most of the neutral hydrogen mass in the Universe at $z\\la 6$ are expected to be hosted by galactic mass dark matter halos~(Wolfe, Gawiser \\& Prochaska~2005). A survey of 21cm intensity fluctuations after reionization would measure the modulation of the cumulative 21cm emission from a large number of galaxies~(Wyithe \\& Loeb~2008; Wyithe, Loeb \\& Geil~2008; Chang et al.~2007). Regarding the measurement of the 21cm PS, this lack of identification of individual galaxies is an advantage, since by not imposing a minimum threshold for detection, such a survey collects all the available signal. This point is discussed in Pen et al.~(2008), where the technique is also demonstrated via measurement of the cross-correlation of galaxies with unresolved 21cm emission in the local Universe. Studying the 21cm PS after (rather than during) reionization offers two advantages. First, it is less contaminated by the Galactic synchrotron foreground, whose brightness temperature scales with redshift as $(1+z)^{2.6}$ ~(Furlanetto, Oh \\& Briggs~2006). Second, because the UV radiation field is nearly uniform after reionization, it should not imprint any large-scale features on the 21cm PS that would mimic the cosmological signatures. In addition, on large spatial scales the 21cm sources are expected to have a linear bias analogous to that inferred from galaxy redshift surveys. Most previous studies of post-reionization 21cm fluctuations have assumed that the 21cm emission traces perturbations in the matter density, and have considered peculiar motions only after a spherical average~(Wyithe \\& Loeb~2008; Pritchard \\& Loeb~2008, Loeb \\& Wyithe~2008). The exception is a recent paper discussing measurements of the dark energy equation of state (Chang et al.~2007). Since the neutral gas resides within collapsed dark matter halos, galaxy bias plays an important role in setting the 21cm fluctuation amplitude. In this paper we derive the 21cm PS after reionization in the context of the formalism that has been developed to calculate it during reionization~(Barkana \\& Loeb~2005; McQuinn et al.~2006; Mao et al.~2008). By framing the derivation this way, the relative merits of cosmological constraints from 21cm surveys at redshifts before and after reionization can be more easily understood. In addition, this formalism provides a framework to describe the possible effect of fluctuations in the ionizing background. We then compute the Alcock-Paczynski effect~(Alcock \\& Paczynski~1979) as an example for the cosmological utility of the post-reionization 21cm PS. In our numerical examples, we adopt the standard set of cosmological parameters ~(Komatsu et al.~2008), with values of $\\Omega_{\\rm b}=0.24$, $\\Omega_{\\rm m}=0.04$ and $\\Omega_Q=0.76$ for the matter, baryon, and dark energy fractional density respectively, and $h=0.73$, for the dimensionless Hubble constant. ", "conclusions": "In this paper we have derived the 21cm power spectrum (PS) following the completion of reionization. Our approach is to derive the PS within the formalism that has been developed to calculate the 21cm PS during reionization~(Barkana \\& Loeb~2005; McQuinn et al.~2006; Mao et al.~2008). By framing the derivation this way we are able to directly compare the relative merits of cosmological constraints from 21cm surveys at redshifts prior to and post reionization. We have derived expressions for the components of the post-reionization 21cm PS that are due to the density field, the ionization field, and their cross-correlation. As is the case prior to reionization, we find that these components contribute to the observed PS in proportions that depend on the angle relative to the line-of-sight along which the power is measured. However, in difference from the situation prior to reionization, we have shown that all components of the 21cm PS are directly proportional to the PS of the underlying matter fluctuations, with a small but predictable modulation on scales below the mean-free-path of ionizing photons. We have derived the form of this modulation, and have shown that its effect on the observed PS will be at less than the 1\\% level for physically plausible astrophysical models of DLA systems. The simplicity of the 21cm PS after reionization stands in contrast to the astrophysical uncertainty during the reionization epoch, where HII regions dominate the 21cm signal. This simplicity will enhance the utility of the 21cm PS after reionization as a cosmological probe (Loeb \\& Wyithe 2008) by removing the need to separate the PS into physical and astrophysical components (Barkana \\& Loeb 2005). To illustrate the utility of the 21cm PS after reionization as a cosmological probe, we have examined the Alcock-Paczynski~(1979) test. Our calculations show that the next generation of low-frequency arrays could measure the angular distortion of the PS to around $\\sim 1\\%$ at $z\\sim3.5$. It has previously been shown that the scale of Baryonic Acoustic Oscillations, which constitutes a standard ruler ~(Blake \\& Glazebrook~2003; Seo \\& Eisenstein~2005; Eisenstein et al.~2005; Padmanabhan et al.~2007), can be also be used to independently probe $H$ and $D_{\\rm A}$ in 21cm PS, and hence to measure the equation of state of the dark energy ~(Wyithe, Loeb \\& Geil~2008; Chang et al.~2007). With the caveat that non-linear evolution of the 21cm PS must be quantitatively understood, our simple analysis indicates that the precision achievable via the Alcock-Paczynski~(1979) test using the 21cm PS after reionization could be better than those available via a 21cm fluctuation measurement of the acoustic scale of baryonic oscillations for a given observing strategy (Shoji, Jeong \\& Komatsu 2009). More generally, our analysis shows that the 21cm PS after reionization shares the same favorable features as a galaxy redshift survey. The advantage of using 21cm fluctuations lies in the fact that individual sources need not be resolved. This would allow PS measurements using 21cm fluctuations to be extended to higher redshifts. \\bigskip \\noindent {\\bf Acknowledgments.} The research was supported by the Australian Research Council (JSBW), by NASA grants NNX08AL43G and LA, by FQXi, and by Harvard University funds (AL). We thank an anonymous referee for correcting an earlier error." }, "0808/0808.0921_arXiv.txt": { "abstract": "We present a high signal-to-noise spectrum of a bright galaxy at $z$ = 4.9 in 14 h of integration on VLT FORS2. This galaxy is extremely bright, $i_{850} = 23.10 \\pm 0.01$, and is strongly-lensed by the foreground massive galaxy cluster Abell 1689 ($z=0.18$). Stellar continuum is seen longward of the Ly$\\alpha$ emission line at $\\sim7100$ \\AA, while intergalactic H~I produces strong absorption shortward of Ly$\\alpha$. Two transmission spikes at $\\sim$6800 \\AA \\ and $\\sim$7040 \\AA \\ are also visible, along with other structures at shorter wavelengths. Although fainter than a QSO, the absence of a strong central ultraviolet flux source in this star forming galaxy enables a measurement of the H~I flux transmission in the intergalactic medium (IGM) in the vicinity of a high redshift object. We find that the effective H I optical depth of the IGM is remarkably high within a large 14 Mpc (physical) region surrounding the galaxy compared to that seen towards QSOs at similar redshifts. Evidently, this high-redshift galaxy is located in a region of space where the amount of H~I is much larger than that seen at similar epochs in the diffuse IGM. We argue that observations of high-redshift galaxies like this one provide unique insights on the nascent stages of baryonic large-scale structures that evolve into the filamentary cosmic web of galaxies and clusters of galaxies observed in the present universe. ", "introduction": "Hydrodynamic simulations tell us that dark matter near the epoch of galaxy formation collapses into an ordered filamentary pattern, the so-called ``cosmic web.\" In turn, this cosmic web is thought to cradle high-redshift galaxies in dense nodes that are opaque to the extragalactic ultraviolet background radiation. Observations of H~I surrounding these high-redshift objects provide information on the likely reservoirs from which galaxies assemble their gas. Hitherto, it has only been possible to measure H~I opacities towards bright Quasi-stellar Objects (QSOs); unfortunately, the high ultraviolet flux from the QSOs ionizes the hydrogen clouds in their vicinity, thereby making the determination of H~I cloud physical conditions unreliable close to the QSO. New techniques and facilities in the past decade have enabled detailed observational studies of the IGM in the early universe ($z > 5$). For example, the spectra of high-redshift QSOs show a plethora of H~I Ly$\\alpha$ absorption lines, the ``Ly$\\alpha$ forest\", as well as lines from a wide variety of heavier elements, all of which provide detailed information about the high-redshift IGM along the line-of-sight towards the QSO. Around the QSOs themselves, however, the ultraviolet flux is so high that it typically ionizes the surrounding gas. Thus QSOs are not ideal for studying the pervasive IGM close to the QSO. Instead, we suggest that bright galaxies may provide better probes for the study of IGM conditions in the proximity of high-redshift objects, as these objects emit fewer ultraviolet photons than QSOs do. The subject of this paper is the Ly$\\alpha$ forest in the spectrum of an unusually bright high-redshift galaxy at $z = 4.866$ (Frye et al. 2002, 2007). The galaxy is situated behind the massive galaxy cluster Abell 1689 ($z = 0.183$), which magnifies the starlight of this background object by a factor of 10.3 by the effect of gravitational lensing (Broadhurst et al. 2005). Multicolor Hubble Space Telescope (HST) imaging of one of the faint lensed images, hereafter designated $A1689\\_7.1$, is shown in Figure 1. We assume a cosmology for this paper of $H_0 = 70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{m,0} = 0.3$, and $\\Omega_{\\Lambda,0} = 0.7$. \\begin{figure} \\includegraphics[viewport=60 -10 250 350,scale=0.7]{f1_small.pdf} \\caption{HST Advanced Camera for Surveys (ACS) $gri$ optical photo of $A1689\\_7.1$ and surrounding region near the center of the galaxy cluster Abell 1689 ($z=0.18$). A single galaxy at $z=4.9$ has been gravitationally-lensed by this foreground galaxy cluster (with member galaxies visible here as yellow spheroid-shaped objects) into a system of three arclets; $A1689\\_7.1$ is the brightest of the three and is shown above and in the inset. This arclet is extremely bright, $i_{775}=23.10 \\pm 0.01$ magnitudes, after being magnified by a factor of $10.3$. $A1689\\_7.1$ is also spatially resolved, with an estimated unlensed angular size of 0.094\\arcsec, corresponding to an intrinsic linear size of only 600 pc. $A1689\\_7.1$ was discovered as a part of a ground-based galaxy redshift survey (Frye et al. 2002, 2007). \\label{fig_image}} \\end{figure} \\begin{figure} \\includegraphics[viewport = 0 0 700 520, scale=0.35]{f2.pdf} \\caption{VLT FORS2 spectrum of $A1689\\_7.1$ at $z=4.866$. The stellar continuum of the galaxy is seen longward of the Ly$\\alpha$ emission line at $\\sim7100$ \\AA, while intergalactic H~I produces strong absorption shortward of Ly$\\alpha$. Two transmission spikes at $\\sim$6800 \\AA \\ and $\\sim$7040 \\AA \\ are also visible, along with other structures at shorter wavelengths. The data enable a measurement of the flux transmission along the line of sight to this star-forming galaxy. The best-fit \\citet{Bruzual:03} model spectrum without corrections for dust extinction and attenuation by the intervening Ly$\\alpha$ forest is overlaid, including a gray region marking our 1$\\sigma$ continuum-placement uncertainties. We select a Chabrier inital mass function (Padova 1994), stellar evolution tracks, and solar metallicity for the model, and give the details regarding the fitting procedure in \\S4. The inset figure shows a theoretical H~I Ly$\\alpha$ profile absorption fit to the data, with horizontal bars delimiting the physical extents of the proximity zone for a QSO (red solid line), and the extent of the first of the five equally-spaced redshift bins in our study over which we measure the H~I opacity (blue-dashed line), respectively. Owing to the lack of a significant proximity zone for our galaxy, a new physical scale is probed close to a high redshift object within a $\\sim14$ physical Mpc radius (blue dashed line).\\label{fig_spec }} \\end{figure} ", "conclusions": "Despite the strong magnification, the intrinsic luminosity of the galaxy is not high; we measure an unabsorbed luminosity of $L_{1400} = 7.7 \\times 10^{28}$ ergs s$^{-1}$ Hz$^{-1}$, and an unlensed $K = 24.5$, two magnitudes fainter than $K^*$ at $z=3$ \\citep{Shapley:01}. Galaxies at $z \\apg 5$ are faint, and hence only a few high quality spectra exist (Price et al. 2007; Kawai et al. 2006; Dow-Hygelund et al. 2007; Franx et al. 1997). Some of these are at $z > 6$ where $\\tau_{GP}^{eff}$ is expected to be high (from the QSO measurements), but the spectrum of the gamma-ray burst GRB060510B at $z = 4.94$ (Price et al. 2007). shows some similar characteristics to the spectrum of $A1689\\_7.1$. The IGM transmission measured over a broad redshift bin of the former is $T = 0.18$, similar to the values observed toward QSOs. However, this broad average smears out the structure in $\\tau_{GP}^{eff}$ as a function of redshift. The GRB060510B spectrum shows evidence of higher than expected GP optical depth extending beyond the region affected by the damped Ly$\\alpha$ absorption at the redshift of the GRB host. It would be interesting to reanalyze this GRB spectrum using smaller redshift bins. H~I opacities were also measured inside the proximity zones of QSOs; although significantly ionized, the proximity zones are found to have neutral hydrogen fractions that exceed theoretical expectations. This indicates that at least some QSOs are also found in regions of gaseous overdensity with large sizes of $15h^{-1}$Mpc \\citep{Guimaraes:07}. Might $A1689\\_7.1$ still be in the process of accreting much of its mass from its overdense surrounding cosmic structure? Theoretical models predict that H~I gas is funneled into young galaxies via large-scale filamentary structures \\citep{Keres:05, Birnboim:03}. We have presented here observational evidence that H~I column densities are higher than expected near one high-redshift galaxy. Based on the large physical size implied by the H~I excess, it is unlikely that this gas will accrete onto a single galaxy. Alternatively, post-starburst galaxies are known to drive high-velocity outflows with velocities of $\\apll 2000$ km s$^{-1}$ in low-ionization stages such as Mg~II \\citep{Tremonti:07}. Given sufficient time, the observed excess of H~I optical depth could be explained by such an outflow; however $A1689\\_7.1$ is too young by a significant factor given the best fit stellar age of 100$\\pm 14$ Myr. On the other hand, QSOs are known to drive outflows with terminal velocities in excess of $10^4$ km s$^{-1}$. $A1689\\_7.1$ does not show obvious signatures of being an AGN, but even if there was a low-metallicity outflow undetected in our spectrum, the velocities would still fall short by more than a factor of ten of that required to explain the H~I excess. As more data become available to study the neutral H~I absorption towards the several recently discovered strongly-lensed LBGs, it may be found to be typical for high-redshift objects, both galaxies {\\it and} QSOs, to be located in regions of space with neutral hydrogen gas fractions significantly larger than that of the pervasive IGM. This re-evaluation of the H~I structure in the densest regions of the universe at these epochs will provide the necessary calibration for modeling the formation and evolution of galaxies in the early universe." }, "0808/0808.2115_arXiv.txt": { "abstract": "Models of galactic chemical evolution (CEMs) show that the shape of the stellar initial mass function (IMF) and other assumptions regarding star formation affect the resultant abundance gradients in models of late-type galaxies. Furthermore, intermediate mass (IM) stars undeniably play an important role in the buildup of nitrogen abundances in galaxies. Here I specifically discuss the nitrogen contribution from IM/AGB stars and how it affects the N/O-gradient. For this purpose I have modelled the chemical evolution of a few nearby disc galaxies using different IMFs and star formation prescriptions. It is demonstrated that N/O-gradients may be used to constrain the nitrogen contribution from IM/AGB-stars. ", "introduction": "Much progress has been made in modelling nucleosynthesis in stars in general, but the origin of nitrogen still remains somewhat uncertain. In the classical picture, the apparent increase in nitrogen production with metallicity, as inferred from spectroscopy of galactic and extragalactic HII-regions, is a manifestation of nitrogen being either primary or secondary (produced from primordial elements or with heavier elements as \"seeds\"). However, reality often turns out to be more complex than our simplest models of it. The origin of nitrogen is probably no exception. During the last few years, reasonably accurate stellar nitrogen abundances have become available also for Population II stars (see, e.g., \\cite{Cayrel04, Ecuvillon04, Israelian04, Spite05}). The rather large nitrogen abundances relative to iron seen in halo stars in the Milky Way suggest that a relatively large fraction of the nitrogen should come from Type II supernovae, especially at low metallicity. However, recent stellar evolution models, including nucleosynthesis, all appear to predict significant nitrogen production in intermediate-mass (IM) stars (\\cite{Marigo01, Izzard04, Gavilan05, Karakas07}). The typical production factor would be of the order $\\sim 10^{-3}$ of the initial stellar mass. If IM-stars of all metallicities produce that much nitrogen, it would require that metal-poor galaxies (or regions within a galaxy) with low N/O-ratios have very young stellar populations, in fact, they have to be so young that IM-stars have not yet made any significant contribution to the interstellar nitrogen abundance of these objects, unless the production of oxygen is much greater at low metallicity. In this work I have considered radial gradients of O/H and N/O in the Milky Way and five other nearby late-type galaxies in order to find additional clues to the origin of nitrogen. The advantages of considering the abundance gradients within galaxies, rather than over-all trends obtained from the global properties of galaxies, are that the same stellar nucleosynthesis prescriptions (yields) must be able to simultaneously reproduce the radial abundance gradients in galaxies with quite different properties (sizes, total masses, gas distributions, etc.) and the very same yields must also be consistent with the fact that galaxies with similar O/H-gradients may have quite different N/O-gradients. Models of chemical evolution (CEMs) may thus be able to provide insight and constraints on stellar yields when several galaxies are modelled and analysed in parallel. ", "conclusions": "Employing a standard IMF (\\cite{Scalo86}) with a high-mass cut-off at $m_{\\rm up}=60/70 M_\\odot$ the O/H-gradients are well-reproduced by numerical CEMs. However, the N/O-gradients cannot, in general, be reproduced simultaneously. This inconsistency can be explained in several ways, e.g., \\begin{enumerate} \\item the production of nitrogen in intermediate-mass stars is strongly correlated with metallicity, \\item the IMF varies over the galactic disc (or, effectively, from galaxy to galaxy) such that the N/O-ratio is affected, \\item the infalling gas that forms the disc is not primordial, but significantly enriched with heavier elements (from Pop. III or Pop. II stars?), \\item or the P-method produces systematic errors. \\end{enumerate} I believe that the explanations (a) and (b) are more likely then the other two. The main reasons are: (1) many ingredients of stellar evolution models that affects the nucleosynthetic yields (e.g., mass loss rates and the efficiency of hot-bottom burning) are still highly uncertain, and (2) the turn-over in the IMF at low stellar masses appears correlated with the Jeans mass (\\cite{vanDokkum08}), (3) significant effects from chemically enriched infall is only expected in the outer parts of the discs, where the abundances are low, and (4) the P-method is in very good agreement with the standard electron-temperature method (\\cite{Pilyugin01a, Pilyugin01b}). Furthermore, the stellar yields and detailed properties of the IMF are probably the greatest uncertainties in CEMs in general. \\begin{figure} \\begin{center} \\includegraphics[width=132mm]{mattsson1_fig4} \\end{center} \\caption{\\label{imf} Variations of the IMF. Changing the slope of a simple power-law IMF ($m_{\\rm up} = 50 M_\\odot, $ $m_{\\rm low}= 0.08 M_\\odot$ are left unchanged) will shift the O/H-gradient up or down (see left panel), but leave $d({\\rm O/H})/d{\\rm R}$ almost unaffected. A Salpeter (1955) IMF ($x = 1.35$) which provides an acceptable fit to the observed N/O-ratios (right panel).} \\end{figure}" }, "0808/0808.4113_arXiv.txt": { "abstract": "The red spectral shape of the visible to near infrared reflectance spectrum of the sharply-edged ring-like disk around the young main sequence star HR\\,4796A was recently interpreted as the presence of tholin-like complex organic materials which are seen in the atmosphere and surface of Titan and the surfaces of icy bodies in the solar system. However, we show in this {\\it Letter} that porous grains comprised of common cosmic dust species (amorphous silicate, amorphous carbon, and water ice) also closely reproduce the observed reflectance spectrum, suggesting that the presence of complex organic materials in the $\\hra$ disk is still not definitive. ", "introduction": "$\\hra$ is a nearby (distance to the Earth $d\\approx 67\\pm 3\\pc$) young main-sequence (MS) star (age $\\approx 8\\pm 3\\myr$; Stauffer et al.\\ 1995) of spectral type A0V (effective temperature $\\Teff\\approx 9500\\K$) with a large infrared (IR) excess which has recently aroused considerable interest. Imaging observations describe dust scattering in the optical (Debes et al.\\ 2008) and near-IR (Augereau et al.\\ 1999; Schneider et al.\\ 1999) and thermal emission at mid-IR (Jayawardhana et al.\\ 1998; Koerner et al.\\ 1998; Telesco et al.\\ 2000; Wahhaj et al.\\ 2005). They reveal a ring-like disk with maximum at $\\simali$70$\\AU$ distance from the central star and $\\simali$17$\\AU$ width that is sharply truncated at the inner and the outer edge. The structure of the HR\\,4796A disk has important implications for planetesimal evolution (Kenyon et al.\\ 1999). Furthermore, possibly existing planets may generate disk asymmetries through gravitational confinement or perturbation (Wyatt et al.\\ 1999) or form rings with sharp edges (Augereau et al.\\ 1999, Klahr \\& Lin 2000, Th\\'ebault \\& Wu 2008). Very recently, Debes et al.\\ (2008) measured a visible to near-IR photometric reflectance spectrum of the dust ring around $\\hra$. To fit the observed spectrum (which is characterized by a steep red slope increasing from $\\lambda \\approx 0.5\\mum$ to 1.6$\\mum$ followed by a flattening of the spectrum at $\\lambda >1.6\\mum$), Debes et al.\\ (2008) argued for the presence of tholin-like organic material in the disk around $\\hra$. Tholin, a complex organic material, was detected as a major constituent of the atmosphere and surface of Titan and the surfaces of icy bodies in the solar system. The detection of tholin in the $\\hra$ disk -- if confirmed -- would imply that these potential basic building blocks of life may be common in extra-solar planetary systems as well (see van Dishoeck 2008 for an overview of organic matter in space). Its young age places HR\\,4796A at a transitional stage between massive gaseous protostellar disks around young pre-MS T-Tauri and Herbig Ae/Be stars ($\\simali$1$\\myr$) and evolved and tenuous debris disks around MS ``Vega-type'' stars ($\\simali$100$\\myr$; see Jura et al.\\ 1993, Chen \\& Kamp 2004). Detecting organic matter in this system will provide valuable information about the formation and evolution of planetary systems. The observed thermal emission of the disk, however, was previously reproduced with a model population of porous grains consisting of the common cosmic dust species amorphous silicate, amorphous carbon, and water ice (Li \\& Lunine 2003a). In this {\\it Letter} we question the existence of tholin as a major dust component in the HR\\,4796A disk and study alternative dust models to reproduce the observed reflectance spectrum. ", "conclusions": "\\label{discussion} Debes et al.\\ (2008) calculated the reflectance spectra of ``astronomical silicates'', water ice, hematite Fe$_2$O$_3$ (which is found on the surface of Mars and responsible for its redness), and UV laser ablated olivine (which has been used to explain the spectral reddening of silicate-rich asteroids due to space weathering; Brunetto et al.\\ 2007). But the scattering spectra of these minerals and ice are too neutral at $\\lambda$\\,$\\sim$\\,0.5--1.6$\\mum$ to match the steep red spectral slope of the reflectance spectrum of the $\\hra$ disk. Debes et al.\\ (2008) therefore resorted to organic materials. They found that tholin or its mixture with other dust species (e.g. water ice or olivine) are able to reproduce the observed red spectral slope. This led them to suggest that ``{\\it the presence of organic material is the most plausible explanation for the observations}'', with a cautionary note that ``... {\\it longer wavelength scattered light observations will further constrain the (tholin-based) grain models, particularly around 3.8--4$\\mum$, where a large absorption feature is seen for different grain sizes of tholins. This would help to directly confirm whether Titan tholins are an adequate proxy for the material in orbit around HR\\,4796A.}'' However, as shown in Figure \\ref{fig:kml_refl}, simple porous dust models consisting of dust species (amorphous silicate, amorphous carbon, water ice) which are commonly considered to dominate in the interstellar medium (ISM), envelopes around evolved stars, and dust disks around young stars closely reproduce the observed reflectance spectrum of the $\\hra$ disk. Our model provides at least a viable alternative to the tholin-based models of Debes et al.\\ (2008). While the tholin organic dust model predicts a strong feature around 3.8--4$\\mum$ characteristic of tholin (Debes et al.\\ 2008), the porous dust model presented here predicts a strong band at $\\simali$3.1$\\mum$, attributed to the O--H stretching mode of water ice. The tholin organics of which the optical constants were adopted by Debes et al.\\ (2008) were made from DC discharge of 90\\% N$_2$ and 10\\% CH$_4$ gas mixture (Khare et al.\\ 1984). They are extremely N-rich and optically very different from amorphous carbon. The dust in the HR\\,4796A disk should be continuously replenished. This is indicated by the low size cutoff $\\amin$ (of a few micrometers) of the dust required to reproduce the reflectance spectrum (see \\S2 and Table \\ref{tab:para}) combined with considerations of dust lifetimes based on the radiation pressure and Poynting-Robertson effects (see Fig.\\,9 of Li \\& Lunine 2003a). The replenishing source would likely arise from collisional cascades of larger bodies like planetesimals, asteroid-like and comet-like bodies. We suggest, with interstellar dust as the building blocks of the parent bodies, it is more reasonable to assume that the dust in the HR\\,4796A disk is composed of amorphous silicate, amorphous carbon, and water ice.\\footnote{% The tholin model may still be viable if the dust in the $\\hra$ disk originates from the surface layers of planetesimals, possibly through excavating collisions (J.H. Debes, private coomunication). In this scenario, the entire planetesimal need not be composed of tholins; they might reside only on the surface where they are created and then be released. Due to significant processing and possible alteration through planetesimal formation, the surface compositions of planetesimals in the $\\hra$ disk may not resemble the ISM composition. Based on what we know from our own Solar System, methane ice and water ice may reside on the surfaces of large planetesimals. These ices are exposed to the stellar UV flux and may ultimately produce tholin-like organic residues. } The model which best fits both the observed reflectance spectrum and the IR emission requires highly porous dust (with $P=0.73$ which corresponds to $P=0.90$ if ice is sublimated; see \\S2). While it is natural to recognize that cold conglomeration of dust grains in molecular clouds can lead to highly porous dust structures,\\footnote{% A porosity in the range of $0.80\\simlt P\\simlt 0.97$ is expected for dust aggregates formed through coagulation as shown both theoretically (Cameron \\& Schneck 1965; Wada et al.\\ 2008) and experimentally (Blum et al.\\ 2006). } at a first glance, it is harder to accept that comparatively more violent collisions between larger bodies would result in such a morphology in the resulting debris. To address this concern, we take the interplanetary dust particles (IDPs) as an analog for the dust in debris disks. The anhydrous chondritic IDPs collected in the stratosphere possibly of cometary origin show a highly porous structure (Brownlee 1987). Love et al.\\ (1994) have measured the densities of $\\simali$150 unmelted chondritic IDPs with diameters of $\\simali$5--15$\\mum$, using grain masses determined from an absolute X-ray analysis technique with a transmission electron microscope and grain volumes determined from scanning electron microscope imaging. They found that these particles have an average density of $\\simali$2.0$\\g\\cm^{-3}$, corresponding to a moderate porosity of $\\simali$0.4. More recently, Joswiak et al.\\ (2007) identified 12 porous cometary IDPs (based on their atmospheric entry velocities) with an average density of $\\simali$1.0$\\g\\cm^{-3}$, corresponding to $P\\approx 0.7$. Much higher porosities ($>$0.9) have been reported for some very fluffy IDPs (e.g. MacKinnon et al.\\ 1987, Rietmeijer 1993), despite that highly porous IDPs are probably too fragile to survive atmospheric entry heating. Low densities are also derived for different groups of meteoroids\\footnote{% By definition, meteoroids are small bodies in the mass range of $\\simali$$10^{-4}$--$10^8\\g$, which orbit the Sun in interplanetary space. The atmospheric trajectories of meteors, i.e. the brightness generated by meteoroids passing through atmosphere, contain information about meteoroid orbits and densities. Meteoroids are accordingly classified into groups with different orbits, structure and composition. The densities given above are valid for between $\\simali$50\\% and 73\\% of cometary meteor observations, depending on the observation method. The same studies show densities of $\\simali$2$\\g\\cm^{-3}$ for the remaining cometary meteoroids, as well as for a significant part of the meteoroids ascribed to asteroids, which is also moderately porous (see Mann 2008). } ascribed to comets: the densities are $\\simali$1.0, 0.75, and 0.27$\\g\\cm^{-3}$, respectively (Ceplecha et al.\\ 1998), corresponding to a porosity of $\\simali$0.7, 0.8, and $>$0.9. It is therefore reasonable to assume that the dust in the $\\hra$ disk has a high porosity (at least for those originated from cometary bodies). We are not sure about the relative contributions to the $\\hra$ disk from asteroid collisions and cometary activity. Note that this is a long-standing problem even for our own solar system (e.g. see Lisse 2002). To conclude, we argue that the presence of tholin-like complex organic materials in the $\\hra$ disk is still not conclusive since the observed red spectral shape of the disk can be closely reproduced by models of porous dust comprised of common cosmic dust species. A more thorough study of the scattered light over a range of scattering angles would further constrain the optical properties of the dust." }, "0808/0808.4055_arXiv.txt": { "abstract": "Based on 2MASS J and Ks photometry for the open star clusters NGC 2383, NGC 2384, Pismis 6, Pismis 8 and using color magnitude diagrams with isochrones fit, we found an age of $\\log (\\mbox{age})$ = 8.3 (200 $\\pm$ 6 Myr) for NGC 2383 and $\\log (\\mbox{age})$ = 6.9 (8 $\\pm$ 6 Myr) for NGC 2384. For Pismis 6 and Pismis 8 we adopted a range of $\\log (\\mbox{age})$ = 6 - 7 (1 - 10 Myr). Because they similar ages, Pismis 6 and Pismis 8 may have been formed in the same Giant Molecular Cloud, and we concluded they are a good candidate for a binary system. In the case of NGC 2383 and NGC 2384, because the big age difference found we conclude that most probably they are born in different environments and as well are not physically connected. ", "introduction": "Open clusters are very important objects in the study of stellar evolution because their members are all of very similar ages and chemical composition. This way the effects of other more subtle variables on the properties of stars are much more easily studied than they are for isolated stars. The total number of open clusters known in our Galaxy is over 1600, (see \"New catalog of optically visible open clusters and candidates\" Dias et al. \\cite{dias1}) of these the only well established double or binary cluster is NGC 869 and NGC 884 (known also as $h + \\chi$ Persei), located at a distance of more than 2 kpc from the Sun. The existence of other possible double clusters has been proposed earlier from Pavloskaya et al. \\cite{pav2}, but not been seriously looked into. Subramaniam et al. \\cite{sub3} examined existing catalogues of open clusters and suggested 18 probable binary open star clusters. The aim of this study is to determine the ages of two probable couples NGC 2383 - NGC 2384 and Pismis 6 - Pismis 8. Our research is based on J and Ks photometry from Two Micron All Sky Survey (2MASS project used two highly-automated 1.3-m telescopes, and provide all-sky photometry in J (1.25 microns), H (1.65 microns), and Ks (2.17 microns) bands). ", "conclusions": "Using 2MASS J and Ks photometry for the open star clusters NGC 2383, NGC 2384, Pismis 6 and Pismis 8, and fitting CMDs with isochrones based on the Geneva models, we found $\\log (\\mbox{age})$ = 8.3 (200 $\\pm$ 6 Myr) for NGC 2383, $\\log (\\mbox{age})$ = 6.9 (8 $\\pm$ 6 Myr) for NGC 2384, and range of $\\log (\\mbox{age})$ = 6 - 7 (1 - 10 Myr) for Pismis 6 and Pismis 8. Pismis 6 and Pismis 8 have similar age and may have been formed in the same GMC, and we conclude that they are a good candidate for binary cluster. In contrast NGC 2383 and NGC 2384 have a big age range between, and may be not formed in the same GMC." }, "0808/0808.3443_arXiv.txt": { "abstract": "We present the results from arcsecond resolution observations of various line transitions at 1.3 mm toward hypercompact HII region G28.20-0.04N. With the SMA data, we have detected and mapped the transitions in the CH$_{3}$CN, CO, $^{13}$CO, SO$_{2}$, OCS, and CH$_{3}$OH molecular lines as well as the radio recombination line H30$\\alpha$. The observations and analysis indicate a hot core associated with G28.20-0.04N. The outflow and possible rotation are detected in this region. ", "introduction": "The scenario of massive star formation remains unclear and has been observationally challenging because of large distances, clustered formation environments, and shorter evolutionary timescales of massive stars. Ultracompact HII (UCHII) regions are considered signposts of massive star formation, but do not represent the earliest stage of massive star forming process (e.g., Churchwell 2002). Hot molecular cores are defined as compact ($\\leq$ 0.1 pc), dense ($\\geq 10^7$ cm$^{-3}$) and warm ($\\geq$ 100 K) molecular cloud cores (Kurtz et al. 2000). The observations from various wavelengths have suggested that hot cores are the sites for massive star formation, and that they represent the early phase of the evolution prior to UCHII regions (Kurtz et al. 2000; Gibb et al. 2000, 2004; Churchwell 2002). So far only twenty or so hot cores from high-resolution observations have been reported in the literature. The observations of hot cores are important for understanding the evolutionary sequence and physical conditions of massive star formation. Hypercompact HII (HCHII) regions have smaller sizes and higher densities compared with UCHII regions, and they probably correspond to a transition phase from hot cores to UCHII regions. The HCHII region G28.20-0.04N at a distance of 5.7 kpc has been observed at radio and millimeter wavelengths (Fish et al. 2003; Sollins et al. 2005; Sewilo et al. 2004, 2008; Keto et al. 2008). Masers, rotation, inflow, and outflow in G28.20-0.04N have been revealed from observations of molecular and radio recombination lines at centimeter wavelengths (Argon et al. 2000; Menten 1991; Sollins et al. 2005; Sewilo et al. 2008). However, the star formation environment and physical conditions of the hot core in this region are still poorly understood. Due to the compact and dense nature of hot cores, high angular resolution observations using molecular lines with high critical densities and excitation temperatures at (sub)millimeter wavelengths are crucial to uncovering the physical conditions and kinematics in G28.24-0.04N. In this letter, we present 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.} observations of molecular lines at 1.3 mm toward the hot core in G28.24-0.04N. ", "conclusions": "" }, "0808/0808.2359_arXiv.txt": { "abstract": "To probe further the possible nature of the unidentified source IGR J17098-3628, we have carried out a detailed analysis of its long-term time variability as monitored by RXTE/ASM, and of its hard X-ray properties as observed by INTEGRAL. INTEGRAL has monitored this sky region over years and significantly detected IGR J17098-3628 only when the source was in this dubbed active state. In particular, at $\\ge$ 20 keV, IBIS/ISGRI caught an outburst in March 2005, lasting for $\\sim$5 days with detection significance of 73$\\sigma$ (20-40 keV) and with the emission at $< $200 keV. The ASM observations reveal that the soft X-ray lightcurve shows a similar outburst to that detected by INTEGRAL, however the peak of the soft X-ray lightcurve either lags, or is preceded by, the hard X-ray ($>$20 keV) outburst by $\\sim$2 days. This resembles the behavior of X-ray novae like XN 1124-683, hence it further suggests a LMXB nature for IGR J17098-3628. While the quality of the ASM data prevents us from drawing any definite conclusions, these discoveries are important clues that, coupled with future observations, will help to resolve the as yet unknown nature of IGR J17098-3628. ", "introduction": "INTEGRAL has detected roughly 500 sources at energies $\\ge$ 20 keV \\citep{boda}. Among them, X-ray binary systems were identified for 32\\% of the times, and another 26\\% remained unidentified. As one of the unidentified sources, IGR J17098-3628 was discovered \\citep{source} at the end of March 2005 by INTEGRAL/IBIS during the private Open Program observation dedicated to the deep view to the Galactic center. The source was located at R.A.(2000) $=17^{h}09^{m}48^{s} $, Dec. $=-36\\degree 28\\arcmin 12\\arcsec $, with an error box of $2\\arcmin$ at 90$\\%$ confidence. The source got a peak flux at $\\sim$ 60 and 95 mCrab for 18-45 and 45-80 keV, respectively. The 20-60 keV flux evolved from $\\sim50$ mCrab at 2005 March 26 00:00, to $\\sim9$ mCrabs at April 3 23:30 (UTC) \\citep{453}. The preliminary spectral analysis \\citep{spe} showed the source spectrum became quite soft along this evolution. Following INTEGRAL, IGR J17098-3628 was observed by RXTE on March 29 of 2005, and was detected at 80 mCrab in the 3--20 keV band, with a hard power-law tail of a spectral index $\\sim$ 2.5. Then the source was assumed a BHC and X-ray Nova (Grebenev et al. 2007). The column density was found to be less than 1$\\times$10$^{22}$ atoms cm$^2$. Later on, Swift observed this source on May 1 of 2005, and refined its location to R.A.\\ (2000) $=17^{\\rm h}09^{\\rm m}45.9^{\\rm s} $, Dec.\\ $=-36\\degree 27\\arcmin 57\\arcsec $, with 90$\\%$ confidence, and an uncertain radius of 5 arcseconds \\citep{kenn}. This made the search possible for the potential counterparts in optical and infrared bands. From the 2MASS ALL-SKY Catalog a possible counterpart was found, J17094612-3627573 \\citep{kong}. However, the most recent optical and infrared observations, made with the 6.5m Magellan-Baade telescope, revealed that 2MASS J17094612-3627573 is in fact composed off several sources \\citep{stee}. Furthermore, Very Large Array (VLA) observations of IGR J17098-3628 were made on March 31, April 5, and May 4, 2005 at 4.86 GHz \\citep{radio}, which led to the discovery of a radio transient at 0.8 arcseconds from the Swift XRT position, and hence regarded as the possible radio counterpart. Finally on July 9, Swift/XRT observed IGR J17098-3628 again and found that the spectrum was fitted by an absorbed disk blackbody model. The column density obtained was (0.89$\\pm$0.02$)\\times$10$^{22}$ atoms cm$^{-2}$ \\citep{kenn2}. In this paper, we report the analysis of all available observations on IGR J17098-3628 carried out by IBIS/ISGRI and JEMX onboard INTEGRAL, and the All Sky Monitor (ASM) onboard RXTE. This allows us to trace the source behavior in X-rays back to 1996 with ASM and back to 2002 with INTEGRAL. We put the outburst in 2005 in this context. ", "conclusions": "IGR J17098-3628 may have stepped into an active phase in the beginning of 2005. INTEGRAL detected a days-long outburst extending up to energies of $\\sim$200 keV. A similar outburst is detected in the soft X-ray (1.5-12 keV) band, however the onset of the outburst lags the hard X-ray outburst by $\\sim$2 days, in a fashion reminiscent of X-ray novae like XN 1124-683 \\citep{nova}. The spectral analysis of the initial stages of the outburst detected by INTEGRAL supports the association with X-ray novae (Grebenev et al. 2007), the cool disk temperature and small inner radius of the accretion disk both suggest that IGR J17098-3628 is a newly discovered black hole system. Current theoretical models of the formation of X-ray novae include the mass transfer instability (MTI) model \\citep{hkl86}, the disk thermal instability (DTI) model (Cannizzo et al. 1982; Faulkner et al. 1983; Meyer \\& Meyer-Hofmeister 1984; Huang \\& Wheeler 1989; Mineshige \\& Wheeler 1989; Ichikawa et al. 1994) Unfortunately, the only way to discriminate between these models relies on detailed differences in the observed soft energy spectrum of the source and so the lack of any detailed X-ray spectra of the IGR J17098-3628 outburst in this energy band prevents us from drawing any conclusions on the X-ray nova mechanism for IGR J17098-3628. That the three episodes of the hard X-ray outburst observed by INTEGRAL have to be described by different models might suggest rather strong time evolution. i.e., the spectral index changed from $\\sim$ 1.7 when rising to roughly 2.6 when decaying suggesting the cooling of the source. Although the time variability of X-ray novae is not well understood theoretically, observations of a variety of X-ray novae not only show that the hard X-ray flux often rises and peaks earlier than the outburst in soft X-rays, but can also show precursor events at hard X-rays several days prior to the main outburst (Chen et al. 1997). While it is not possible to associate the outburst in IGR J17098-3628 with any particular X-ray nova mechanism spectrally, it is clear that the typical accretion scenario for the low mass X-ray binaries holding a black hole is not capable of reproducing the observed lag between the hard and soft X-ray outbursts. In this scenario, the soft X-rays are produced by the accretion disk, while the hard X-rays are produced by comptonization of soft seed photons in the inner part of the accretion disk \\citep{disalvo,stella}. Clearly, large changes in the hard X-ray flux can, at best, be coeval with large changes in the soft X-ray seed photons and cannot precede variations of the soft X-ray photons that the process of comptonization relies on to produce the hard X-ray tail. One potential mechanism that may reproduce the observed lag is for an instability in the disk to trigger the formation of a temporary jet, which produces hard X-rays via inverse comptonization, in advance of any change in the flux of soft X-rays from the disk. A more straightforward scenario in understanding this different time behaviour between the hard and the soft X-rays may lie in the tie-up with the structure of the accretion flow. The precursor event can be explained as the transition from the \"hard\" state to \"soft\" state in the BHC system. The accretion time in the inner parts of the disk is much smaller than the duration of X-ray nova's outburst therefore changes in the spectral state of these sources are connected with general changes in structure of the accretion flow rather than in changes in number of soft seed photons for Comptonization. Observations show that all BH systems have the hard Comptonized spectrum at low accretion rates and the soft DBB(disc blackbody) spectrum (with a possible weak hard tail) at high accretion rates. This may be connected with approaching of the inner radius of a cold accretion disk to the radius of marginally stable orbit when the accretion rate increases. The observed lag between outbursts in hard X-rays and soft X-ray may be directly connected with this observed dependence: the accretion rate is small at the initial stage of the outburst thus the spectrum is hard and the outburst is observed in hard X-rays, but later the accretion rate rises and the spectrum becomes soft thus the outburst is observed in soft X-rays. Although no clear statement can be made regarding their nature, the discovery the lag/precursor event revealed by the concurrent RXTE ASM and INTEGRAL observations in 2005, is important piece of the IGR J17098-3628 puzzle that will help to resolve the as yet unknown nature of this perplexing source." }, "0808/0808.0809_arXiv.txt": { "abstract": "People usually smile when astrophysicists assert that we are {\\it sons of the stars}, but human life confirms this sentence: about 65\\% of the mass of our body is made up of oxygen, carbon occurs in all organic life and is the basis of organic chemistry, nitrogen is an essential part of amino acids and nucleic acids, calcium is a major component of our bones. Moreover, phosphorus plays a major role in biological molecules such as DNA and RNA (where the chemical codes of life is written) and our blood carries oxygen to tissues by means of the hemoglobin (an iron pigment of red blood cells). All these elements have been created in stars. I just list some examples related to human body, but also common element such as aluminum, nickel, gold, silver and lead come from a pristine generation of stars. The abundances in the Solar System are in fact due to the mixing of material ejected from stars that polluted the Universe in different epochs before the Sun formation, occurred about 5 billion years ago, after the gravitational contraction of the proto-solar cloud. Low mass AGB stars (1$<$$M$/M$_{\\odot}$$<$3) are among the most important polluters of the Milky Way, because of the strong winds eroding their chemically enriched envelopes. They are responsible for the nucleosynthesis of the main component of the cosmic s-elements. ", "introduction": "The majority of the isotopes heavier than iron (A$\\ge$56) are synthesized by neutron capture processes. The observed heavy elements distribution shows the presence of two main components, correlated to different nucleosynthetic processes: the s (slow) process and the r (rapid) process (on the basis of the definitions given by \\cite{b2hf} in their pioneering work). The r process requires high neutron densities, and it is believed to occur during explosive phases of stellar evolution (Novae, SuperNovae and/or X-rays binaries). The s process is characterized by a slow neutron capture with respect to the corresponding $\\beta$ decay: stable isotopes capture neutrons, while the radioactive ones decay ($\\beta^-$ or $\\beta^+$) or capture a free electron. These isotopes are mainly created in the Thermally Pulsing AGB (TP-AGB) phase of low mass stars (1 $\\leq$ M/M$_{\\odot}$ $<$ 4), where freshly synthesized elements are carried out to the surface by means of a recurrent mechanism called Third Dredge Up (TDU) (see \\cite{stra06} and references therein). \\begin{figure}[tb] \\centering \\includegraphics[width=8.5cm]{mcore.ps} \\caption{Evolution of the positions in mass of the inner border of (top to bottom): convective envelope, H-burning shell and most energetic mesh of the He-burning shell, during the Thermally Pulsing AGB phase of a model with initial mass M=2M$_\\odot$ and solar metallicity..} \\label{mcore} \\end{figure} In this phase the stellar structure consists of a partially degenerate carbon-oxygen core, an He shell separated from an H shell by the He-intershell region and by a convective envelope (see Fig.\\ref{mcore}). The energy required to supply the surface irradiation is mainly provided by the H burning shell, located just below the inner border of the convective envelope. This situation is recurrently interrupted by the growing up of thermonuclear runaways, driven by violent He-burning ignitions. As a consequence of a Thermal Pulse (TP), the region between the two shells (He-intershell) becomes unstable against convection (for a short period), the external layers expand and, later on, the H shell burning temporarily dies down. In the He-intershell, He is partially converted into carbon. During the AGB phase, main neutron sources are the $^{13}$C($\\alpha$,n)$^{16}$O reaction, active in radiative layers during the interpulse period \\cite{stra95}, and the $^{22}$Ne($\\alpha$,n)$^{25}$Mg reaction, marginally activated within the convective shell originated by the TP. In order to obtain a sufficient amount of $^{13}$C for the activation of the s-process, a diffusion of protons from the H-rich envelope into the $^{12}$C-rich radiative zone is needed: the diffused proton are captured from the abundant carbon via the $^{12}$C(p,$\\gamma$)$^{13}$N($\\beta^-$)$^{13}$C nuclear chain, leading to the formation of a tiny $^{13}$C-pocket. ", "conclusions": "The present work demonstrates that, nowadays, the computational power allows the coupling between a stellar evolutionary code and a full nuclear network. A different treatment of the internal border of the convective envelope with respect to a bare Schwarzschild criterion allows the formation of the so-called $^{13}$C pocket. For the first time in the literature, it has been possible to directly compare theoretical models and observational data both from a physical (luminosities, surface temperatures) and a chemical (light elements and heavy elements abundances) point of view . Moreover, we furnish a uniform set of yields (from hydrogen to lead) at different metallicities. The importance of the adopted mass loss rate and of molecules contribution to opacity has been pointed out. While the first problem is still under analysis, the second one has been solved by using opacity tables with enhanced carbon and nitrogen abundances. Finally, we have addressed the importance of very low metallicity AGB models, by describing first preliminary (and promising) results." }, "0808/0808.0038_arXiv.txt": { "abstract": "We present a sample of 16 radio galaxies, each of which is characterized by a wide, elongated emission gap with fairly sharp and straight edges between the two radio lobes. This particular subset of the ``superdisk'' radio galaxies is chosen because of a highly asymmetric location of the host elliptical galaxy relative to the gap's central axis. In addition to posing a considerable challenge to the existing models, such a morphology also means that the two jets traverse highly unequal distances through the superdisk material. One thus has a possibility to directly investigate if the marked asymmetry between the two jets' interaction with the (much denser) ambient medium, during their propagation, has a significant import for the brightness of the hotspot forming near each jet's extremity. We also propose a new explanation for the formation of superdisks through the merger of a smaller elliptical galaxy with the massive host, in which the gas attached to the infalling galaxy deposits its angular momentum into the host's circumgalactic gas, thereby causing it to flatten into a fat pancake, or superdisk. The asymmetric location of the host galaxy can be assisted by the kick imparted to it during the merger. We also suggest a physical link between these radio galaxies and those with {\\bf X}-shaped and {\\bf Z}-symmetric radio lobes, commonly believed to arise from mergers of two galactic nuclei, each harboring a supermassive black hole. ", "introduction": "A few years ago we noted that a fraction of extragalactic double radio sources exhibit radio morphologies and other properties that led us to infer that they required gaseous ``superdisks\" extending tens of kiloparsecs in both diameter and thickness \\citep[hereafter GKW00]{gkw00}. The key morphological property shared by these radio galaxies is the sharp, quasi-linear edges of the radio lobes on the sides facing the central elliptical host galaxy \\citep{gkw96,gkna}, as seen in at least a dozen radio galaxies at low to moderate redshifts. We showed that such superdisks (SDs) can provide consistent, alternative explanations for some key correlations found among the parameters of extragalactic radio sources, such as the Laing-Garrington effect \\citep[e.g.,][]{garr88,lain} and the correlated radio-optical asymmetries \\citep[e.g.,][]{mcca}. Furthermore, in high-$z$ RGs the strong tendency for redshifted component of the diffuse, quiescent Ly$\\alpha$ emission to be on the side of the brighter hot spot has been explained by postulating that the redshifted Ly$\\alpha$ emission originates from the region of the radio lobe on the near side of the nucleus, and is thus subject to much less dilution due to the intervening dust \\citep[e.g.,][]{hump,gkw05}. As discussed in GKW00, these Ly$\\alpha$ RGs provide optical evidence for high-$z$ SDs. Because the sharp-edged morphology would only be noticed when the radio jets are oriented close to the plane of the sky, SDs are subject to a strong negative selection effect, and GKW00 argued that even though the observed cases are few, the phenomenon may not be so uncommon. In this paper we present an enlarged sample of radio galaxies (RGs) exhibiting SDs and discuss, in particular, their properties related to structural asymmetry. The sharp emission gaps seen in double radio sources are most commonly attributed to a blocking of the backflowing radio plasma in the lobes by a denser thermal plasma associated with the parent elliptical galaxy \\citep[e.g.,][]{leahw,wiitg,blac}. Such backflows are to be expected and simulations indicate that they will indeed be diverted by galactic ISM \\citep[e.g.,][]{wiitn}. This approach was an alternative to the original proposal by \\citet{sche} according to which the central emission gap arose from the pinching of the inner parts of the lobes by the higher gas pressure of the host galaxy. Both of these mechanisms, however, seem incompatible with cases where the edges of the gaps are long and straight; they seem particularly incapable of explaining the highly asymmetric SDs where the host elliptical is found almost at one edge of the radio emission gap. A third possible explanation for large gaps has recently been put forward by \\citet{gerg}; they suggest that SDs could be carved out as two galaxy cores containing supermassive black holes (SMBHs) merge and the associated pair of jets undergoes a rapid precession during the later stages of the merger. While this novel process might indeed create a gap, it too may have difficulty in explaining the sharp inner edges of the lobes, particularly those of the highly asymmetric SDs we are exploring here. Initially we proposed that the SD is primarily made of the interstellar medium bound to the RG itself, perhaps originating from the cool gas belonging to gas-rich disk galaxies previously captured by the giant elliptical host of the powerful RG \\citep[e.g.,][]{stat}. We argued that the tidal stretching and heating of that gas during the capture was sufficient to produce very large fat pancakes \\citep{gkna,gkw00}. Since in some extreme cases (e.g., 0114$-$476; Table 1) SDs were found to have widths running into several hundred kiloparsecs, an alternative possibility was also considered in GKW00, according to which at least some SDs trace the gaseous filaments of the ``cosmic web\". In a recent paper, we have attempted to explain the occurrence of SDs in high-$z$ RGs which have yet to acquire a significant circumgalactic medium (CGM) or reside in an intracluster medium (ICM) \\citep[ICM;][]{gkwj}. We argued that the material forming the SD in such RGs likely arises from the nuclear wind which is believed to be associated with the AGN activity \\citep[e.g.,][]{soke}. We investigated the conditions under which this wind material would be squeezed between the two radio lobes into a pancake shaped SD. It was shown that for a wide range of reasonable wind and jet parameters those jets launched within a few tens of Myr subsequent to the initiation of wind production could quickly catch up to the wind-blown bubble and then the lobes growing outside the bubble could indeed squeeze the bubble in such a way as to produce SDs. Such a squeezing, however, is unlikely in $z < 1$ RGs, which are usually surrounded by a significant amount of X-ray emitting CGM with a pressure at least that of the radio lobes \\citep[e.g.,][]{cros}. Therefore, another mechanism seems to be required at least for the SDs in low $z$ RGs. In \\S 2 we tabulate a significant number of additional RGs with highly asymmetric SDs. We also briefly discuss some interesting morphological properties of these asymmetric SDs. We then propose in \\S 3 a new mechanism based on a merger of an elliptical galaxy, also rich in hot gas, with the massive host elliptical. In this scenario the orbital angular momentum of the captured galaxy's gas will be gradually transferred to the CGM gas of the host galaxy, thus transforming its quasi-spherical CGM halo into a fat pancake. This scenario ties in well with our explanation for the {\\bf Z}-symmetries seen in some {\\bf X}-shaped RGs (XRGs) \\citep[hereafter GKBW03]{gkbw}. Conclusions are summarized in \\S 4. ", "conclusions": "We have highlighted a particularly interesting and intriguing class of RGs which not only shows an SD type morphology but also the host galaxy is seen close to the edge of the SD and is thus grossly offset from the midplane of the SD. This morphological asymmetry poses a serious difficulty for existing SD models, particularly when the superdisk is found in RGs at $z \\lesssim 1$, where a significant amount of hot gas surrounding the host is usually present. We have shown that the two radio lobes extend significantly more symmetrically about the SD midplane than they extend relative to the host galaxy. We have proposed a scenario whereby gaseous SDs acquire their form through the injection of angular momentum from the gas belonging to a galaxy whose eventual merger with the massive elliptical host triggers the jets responsible for the RG. The gradual transfer of the angular momentum into the hot CGM of the host elliptical flattens its CGM into a fat-pancake shaped superdisk, which can account for the strip-like emission gaps observed between the radio lobes of such RGs. In this picture, the grossly off-centered location of the host (relative to the central plane of the SD) is relatively straightforward to understand in terms of the motion of the host galaxy during the period of jet activity, partly assisted by a linear momentum kick imparted by the merged galaxy. A corollary to this picture is that the SD material (which appears to be ``docking the tails'' at least in these RGs \\citep[cf.,][]{jenk} is not gravitationally bound to the radio-loud elliptical. Interestingly, this scenario also provides a possible physical link between the SDs and the XRGs that are widely believed to manifest mergers of two galaxies containing SMBHs. The merger is likely to launch a pair of jets and if the host had already been recently active, a double-double radio structure can arise, as seen in three RGs in the present sample (0035$+$130, 0114$-$476, and 1155$+$256). It is interesting that for between 10 and 13 of these 16 RGs, the brighter of the two outer hotspots is associated with the radio lobe adjoining the host galaxy. This indicates a diminishing of a jet's radio emitting potential following a long passage through the denser thermal material associated with the SD, as compared to a jet that mainly traverses the (light) relativistic plasma within a radio lobe. Perhaps a striking manifestation of this is seen in the RGs 0106$+$729 and 1939$+$605 where the radio jet appears to turn dissipative after propagating through the SD. No such dissipation is observed, at least in the present sample, for the jets that propagate mainly through the radio lobe. This calls for a follow-up study once a significantly larger sample of highly asymmetric superdisks becomes available. We intend to investigate the realm of applicability of this model through simulations of the merger of two massive ellipticals with substantial circumgalactic gas." }, "0808/0808.2384_arXiv.txt": { "abstract": "We report the discovery of ripple-like X-ray surface brightness oscillations in the core of the Centaurus cluster of galaxies, found with 200~ks of \\emph{Chandra} observations. The features are between 3 to 5 per cent variations in surface brightness with a wavelength of around 9~kpc. If, as has been conjectured for the Perseus cluster, these are sound waves generated by the repetitive inflation of central radio bubbles, they represent around $5\\times 10^{42}\\ergps$ of spherical sound-wave power at a radius of 30~kpc. The period of the waves would be $10^7$~yr. If their power is dissipated in the core of the cluster, it would balance much of the radiative cooling by X-ray emission, which is around $1.3 \\times 10^{43} \\ergps$ within the inner 30~kpc. The power of the sound waves would be a factor of four smaller that the heating power of the central radio bubbles, which means that energy is converted into sound waves efficiently. ", "introduction": "Ripples in the X-ray surface brightness of galaxy clusters were first discovered in the Perseus cluster of galaxies \\citep{FabianPer03}. As the surface brightness is roughly proportional to the density of the intracluster medium (ICM) squared, these are density ripples. \\cite{FabianPer03} explained these variations as sound waves generated by the inflation of the bubbles of relativistic plasma in the core of the cluster by the central active nucleus. Providing that these sound waves can be dissipated as they travel \\citep{FabianReynolds05}, they have the ability to transport energy from the central nucleus to the core of the cluster. If they carry significant quantities of energy, they may provide the distributed heating required \\citep{Voigt04} to prevent large quantities of the ICM from cooling \\citep{PetersonFabian06,McNamaraNulsen07}. Subsequent theoretical work \\citep{Ruszkowski04,Sijacki06} have shown that sound waves may be able to help heat the cores of clusters in a gentle distributed fashion. A deep 900~ks observation of Perseus showed the ripples in exquisite detail enabling their amplitude to be measured \\citep{FabianPer06}. Subsequently, we used the amplitude of the ripples to calculate the amount of energy propagated in them if they are sound waves \\citep{SandersPer07}. They would carry enough energy in Perseus to combat a significant fraction of the cooling rate, and appear to decline in power with radius. The Perseus observations also reveal thick spherical shells around each inner bubble at a higher gas density and pressure than their surroundings. These high pressure shells have a sharp outer edge where the density abruptly jumps by about 30 per cent, and is interpreted as a weak shock. We have assumed that such high pressure regions propagate outward to become the ripples. A problem with the weak shock interpretation is that there is no temperature jump coincident with the density one. The observed limit on any jump in temperature is inconsistent with an adiabatic shock. The region has been studied and discussed in detail by \\cite{Graham08Per}. It is noted there that the Northern optical filaments of cold gas stop at the shock front, so cold gas is likely being mixed with the hot gas there due to turbulence generated at that front. This could explain a reduction in temperature of the post shock gas. Sound waves or ripples have not been detected in other systems. However, they may still be an important contribution to the heating processes in clusters as they could be difficult to detect. They were not found until 200~ks of \\emph{Chandra} time of the Perseus cluster was obtained. Perseus is the X-ray brightest galaxy cluster by a factor of 1.7 (in the 2-10~keV band; \\citealt{Edge90}). Clearly the detectability of ripples will vary with cluster surface brightness and observation time. It will also vary with ripple amplitude and wavelength and with cluster properties. These effects have been examined numerically by \\cite{Graham08}, finding that it is difficult to observe these features with current instruments in clusters other than Perseus. The detectability of ripples, however, depends on unknown factors, such as the wavelength and amplitude of the ripples. These factors can be estimated with a large degree of uncertainty from the properties of the cluster, assuming that they are responsible for heating the cluster core and have a wavelength similar to the bubble size. Here we report on some ripples which we find in a 200~ks \\emph{Chandra} observation of the Centaurus cluster of galaxies. These data were examined previously by \\cite{Fabian05} and \\cite{SandersEnrich06}, examining the thermal distribution of gas in the cluster core and its metallicity. The Centaurus cluster of galaxies is a nearby ($z=0.0104$; \\citealt{LuceyCurrieDickens86a}) X-ray bright galaxy cluster ($L_{X,2-10\\keV} = 2.9 \\times 10^{43} \\ergps$; \\citealt{Edge90}). It contains a low radiative efficiency nucleus \\citep{Taylor06} and distinct bubbles of relativistic plasma \\citep{Taylor02} displacing the intracluster medium, as seen by \\emph{Chandra} \\citep{SandersCent02}. We assume $H_0 = 70 \\kmpspMpc$, which translates into scale of 213~pc per arcsec. ", "conclusions": "We detect surface brightness deviations in a \\emph{Chandra} image of the Centaurus cluster of galaxies. These features are between 3 and 5 per cent in size. Assuming that they are sound waves, the wave power is approximately $5\\times 10^{42}\\ergps$ in the inner 30~kpc, with a period of $10^7$ yr and a wavelength of 9~kpc. The power is close to the X-ray luminosity of this central region. It is therefore energetically possible that sound waves are the mechanism by which the power of the central black hole is transmitted quasi-isotropically through the cluster. Provided that the power in the waves is gradually dissipated as heat, they provide the means by which radiative cooling is balanced by heating in the inner cluster core." }, "0808/0808.2667_arXiv.txt": { "abstract": "We present a systematic study of line widths in the [\\ion{O}{3}]$\\lambda$5007 and H$\\alpha$ lines for a sample of 86 planetary nebulae in the Milky Way bulge based upon spectroscopy obtained at the \\facility{Observatorio Astron\\'omico Nacional in the Sierra San Pedro M\\'artir (OAN-SPM)} using the Manchester Echelle Spectrograph. The planetary nebulae were selected with the intention of simulating samples of bright extragalactic planetary nebulae. We separate the planetary nebulae into two samples containing cooler and hotter central stars, defined by the absence or presence, respectively, of the \\ion{He}{2}\\,$\\lambda$6560 line in the H$\\alpha$ spectra. This division separates samples of younger and more evolved planetary nebulae. The sample of planetary nebulae with hotter central stars has systematically larger line widths, larger radii, lower electron densities, and lower H$\\beta$ luminosities. The distributions of these parameters in the two samples all differ at significance levels exceeding 99\\%. These differences are all in agreement with the expectations from hydrodynamical models, but for the first time confirmed for a homogeneous and statistically significant sample of galactic planetary nebulae. We interpret these differences as evidence for the acceleration of the nebular shells during the early evolution of these intrinsically bright planetary nebulae. As is the case for planetary nebulae in the Magellanic Clouds, the acceleration of the nebular shells appears to be the direct result of the evolution of the central stars. ", "introduction": "The interacting stellar winds model of \\citet{kwoketal1978} provides a general framework for understanding planetary nebulae. Within this framework, a variety of theoretical and numerical studies of the hydrodynamics have been undertaken to understand and predict the kinematic properties of planetary nebulae. Our modern view of the kinematics of planetary nebulae was established in the early 1990's \\citep[e.g.,][]{kahnwest1985, breitschwerdtkahn1990, kahnbreitschwerdt1990, mellema1994}. These studies showed that the AGB envelope is accelerated in two phases. The first phase is a result of the shock wave initiated by the ionization front that sweeps through the AGB envelope as the central star's temperature increases. Then, as the central star's wind energy increases, it drives a pressure-driven central bubble that accelerates the AGB envelope through ram pressure. More recent numerical work that attempts to include more realistic AGB evolution confirms these basic results \\citep[e.g.,][]{villaveretal2002, perinottoetal2004}. Eventually, the central star's wind ceases and the inner part of the AGB envelope backfills towards the central star while the outer part maintains its momentum-driven evolution \\citep[e.g.,][]{garciaseguraetal2006}. These models have been used extensively to interpret observational studies of the kinematics of many individual planetary nebulae. Unfortunately, to date, systematic, homogeneous studies of planetary nebula populations to compare with these theoretical efforts are scarce. \\citet{heap1993} found a correlation between nebular expansion velocity and the terminal velocity of the central star wind. More recently, \\citet{medinaetal2006} claimed a correlation between expansion velocity and age indicators such as density and central star temperature. However, both of these studies suffer from heterogeneity, at least regarding object selection. Even regarding planetary nebulae with Wolf-Rayet (WR) central stars, the precise role that the winds from the central stars play in the kinematics of the nebular shells is not entirely clear \\citep{gesickietal2006, medinaetal2006}. The studies of the kinematics of planetary nebulae in the Magellanic Clouds are perhaps the most systematic \\citep{dopitaetal1985, dopitaetal1988}. These studies find correlations between expansion velocity, excitation class, and nebular density, suggesting that the properties of the central star dominate the nebular evolution. As they point out, however, their results are not general since they have been found only for a particular population of (bright) planetary nebulae. Therefore, there remains a need for a coherent population study in a different galactic environment. Here, we present a study of the kinematics for a large sample of planetary nebulae in the bulge of the Milky Way (Bulge). These objects were selected in a homogeneous way, with the hope of simulating populations of bright extragalactic planetary nebulae in bulge-like systems (bulges of spiral galaxies, dwarf spheroidals, and elliptical galaxies). The observations were all made with the same instrument and methodology. Likewise, the data reduction and analysis is homogeneous and entirely independent of model parameters. Details are given in Section 2. We find that the expansion properties do indeed depend upon the evolutionary stage of the central star, with planetary nebulae hosting hotter central stars having larger expansion velocities (Section 3). We also find that other parameters that should depend upon age, the nebular size, density, and H$\\beta$ luminosity, also differ significantly between the two groups (Section 4). We compare these properties with theoretical models and find generally good agreement (Section 5). We present our conclusions in Section 6. ", "conclusions": "We have obtained kinematic data for a large sample of planetary nebulae in the Milky Way bulge, selected with the goal of simulating samples of bright extragalactic planetary nebulae in bulge-like environments. For most of the sample, our criteria also included a reddening-corrected H$\\beta$ flux exceeding $\\log F(\\mathrm H\\beta)>-12.0$\\,dex and [\\ion{O}{3}]$\\lambda 5007/\\mathrm H\\beta > 6$, in addition to the usual criterion of projected proximity to the Galactic center. We measure line widths for H$\\alpha$ and [\\ion{O}{3}]$\\lambda 5007$. Our sample is the largest that has been selected and observed in a homogeneous way. We have also analyzed it in a completely model-independent manner. For half of the sample, the \\ion{He}{2}$\\lambda 6560$ line is observed in the H$\\alpha$ spectra. Since this line appears only for hotter central stars, it allows us to divide our sample according to the evolutionary stage of the central star, with the more evolved objects found in the subsample containing the hotter central stars. The kinematics of the two samples are significantly different, with the sample containing the hotter central stars expanding faster. We also compare the diameters, electron densities (from [\\ion{S}{2}]$\\lambda\\lambda 6716,6731$), and H$\\beta$ luminosities for the two subsamples. In all cases, there are statistically significant differences. The planetary nebulae hosting the hotter central stars are larger, less dense, and less luminous than their counterparts with cooler central stars. All of these differences exceed a statistical significance of 99\\%. Also, all of these differences are compatible with the results of hydrodynamical models \\citep[e.g.,][]{mellema1994, villaveretal2002, perinottoetal2004}. Our primary conclusion is that we have clearly observed the acceleration of the nebular shells in planetary nebulae in the bulge of the Milky Way and that this occurs during the early evolution of their central stars. Our findings, based upon a large, homogeneous sample, constitute the first unequivocal evidence that the nebular kinematics depend upon the evolutionary state of the central star for planetary nebulae in the Milky Way and are far clearer and more convincing than any previous results. Hence, there is now very clear evidence that this dependence occurs in at least two environments: the Magellanic Clouds and the bulge of the Milky Way. Given the differences in the stellar populations in these two environments, it is likely that the dependence of the nebular kinematics on the evolutionary state of central star is more general and that it occurs in all environments." }, "0808/0808.3985_arXiv.txt": { "abstract": "We present {\\sl Suzaku} spectra of X-ray emission in the fields just off the LMC~X--3 sight line. \\ovii, \\oviii, and \\neix\\ % emission lines are clearly detected, suggesting the presence of an optically thin thermal plasma with an average temperature of 2.4 $\\times 10^6$ K. This temperature is significantly higher than that inferred from existing X-ray absorption line data obtained with \\chandra grating observations of \\xs, strongly suggesting that the gas is not isothermal. We then jointly analyze these data to characterize the spatial and temperature distributions of the gas. Assuming a vertical exponential Galactic disk model, we estimate the gas temperature and density at the Galactic plane and their scale heights as $3.6(2.9, 4.7)\\times10^6$ K and $1.4(0.3, 3.4)\\times10^{-3}~{\\rm cm^{-3}}$ and $1.4(0.2, 5.2)$ kpc and $2.8(1.0, 6.4)$ kpc, respectively. This characterization can account for all the \\ovi\\ line absorption, as observed in a \\fuse\\ spectrum of LMC~X--3, but only predicts less than one tenth of the \\ovi\\ line emission intensity typically detected at high Galactic latitudes. The bulk of the \\ovi\\ emission most likely arises at interfaces between cool and hot gases. ", "introduction": "\\label{sec:intro} The Galactic diffuse hot gas at temperatures $\\sim10^6$ K can be effectively probed via its emission and absorption features in X-ray and far-ultraviolet (UV) wavelength bands. The previous X-ray emission investigations were largely based on the broadband X-ray background data, e.g., {\\sl ROSAT ALL Sky Survey} (RASS; Snowden \\etal 1997). A high spectral resolution X-ray calorimeter abroad a sounding rocket, though providing little spatial resolution, clearly detected the \\ovii\\ and \\oviii\\ emission lines, confirming that much of the soft X-ray background (SXB) emission is thermal in origin \\citep{mcc02}. Recently, several groups have attempted to study the background emission with X-ray CCDs aboard {\\sl Suzaku X-ray Observatory}, which, compared to those on {\\sl XMM-Newton} and {\\sl Chandra X-ray observatories}, have a significantly improved spectral resolution and a low instrument background (e.g., Smith \\etal 2007; Henley \\& Shelton 2007). These later X-ray observatories, however, also carry the high resolution grating instruments that allow for the detection of the X-ray absorption lines (e.g., from \\ovii, \\oviii, and \\neix\\ K$\\alpha$ transitions) by diffuse hot gas in and around the Galaxy. Indeed, such absorption lines are detected in grating spectra of nearly all Galactic and extragalactic sources as long as the spectral signal-to-noise ratio is high enough (e.g., Futamoto \\etal 2004; Yao \\& Wang 2005; Fang \\etal 2006; Bregman \\& Llyoid-Davis 2007). These X-ray absorption lines trace gas over a broad temperature range of $\\sim 10^{5.5}-10^{6.5}$~K. Gas at temperatures $\\lsim 10^{5.5}$~K may be traced more sensitively in far-UV, via the detection of the \\ovi\\ lines at $\\lambda\\lambda$1031.96 and 1037.62. Extensive observations of the lines in absorption have been carried out with {\\sl Copernicus} and {\\sl Far Ultraviolet Spectroscopy Explorer} ({\\sl FUSE}; Jenkins 1978; Savage \\etal 2003; Bowen \\etal 2008). In addition, the \\ovi\\ lines have also been detected in emission with {\\sl FUSE}, although the sky is only sampled at various Galactic latitudes (e.g., Shelton \\etal 2001; Otte \\& Dixon 2006). The intensity of the $\\lambda$1031.96 line, for example, ranges from 1800 to 9100 LU (line unit; 1 ${\\rm photon~cm^{-2}~s^{-1}~sr^{-1}}$). However, the interpretation of the \\ovi\\ line(s) alone is not straight forward, because in the collisional ionization equilibrium (CIE) state, the \\ovi\\ population sharply peaks at the intermediate temperature $\\sim 10^{5.5}$~K where gas cools very efficiently \\citep{sut93}. Thus such \\ovi-bearing gas is expected to be rare and may preferentially reside at interfaces between cool gas clouds and thermally more stable hot gas. But this latter hot gas, which should be more abundant, distributed more widely, and effectively traced by the X-ray \\ovii\\ line, could contribute significantly to the \\ovi\\ line as well. Currently, little is known about the relative contributions from these two origins to the \\ovi, either in absorption or in emission. Clearly, a combined analysis of the X-ray and far-UV lines, in both emission and absorption, will be the most beneficial. While an absorption line is proportional to the total column density of the gas integrated along a line of sight, an emission line depends on the emission measure (EM) of the gas. Furthermore, for a gas in the CIE state, both the ionic column density and the EM depend on the gas temperature, but in different manners (Fig.~\\ref{fig:absVSemi}). Therefore a joint analysis of multiple emission/absorption lines will enable us to constrain not only the temperature and its distribution but also the size and density of the intervening gas (e.g., Shull \\& Slavin 1994). When transitions from multiple elements (e.g., O and Ne) are detected, their relative abundances can also be estimated (Yao \\& Wang 2006). Such a joint analysis has been tentatively applied to several sources (e.g., Yao \\& Wang 2007; Shelton \\etal 2007); but none of these sources has high resolution emission and absorption data available in both X-ray and far-UV wavelength bands. \\begin{figure} \\plotone{f1.eps} \\caption{Ionization fraction (a) and the emissivity (b) of oxygen ions as a function of temperature for a gas in the collisional ionization equilibrium state \\citep{sut93}. The emissivity is a summation of the doublet transitions at 1031.96 \\AA\\ and 1037.62 \\AA\\ (\\ovi), triplet transitions at 22.10 \\AA, 21.80 \\AA, and 21.60 \\AA\\ (\\ovii), and K$\\alpha$ plus K$\\beta$ transitions at 18.97 \\AA\\ and 16.0 \\AA\\ (\\oviii). The emissivity of \\ovi\\ is scaled down by a factor of 1000 for demonstration purpose. \\label{fig:absVSemi} } \\end{figure} In this paper, we report our investigation of the hot gas along the sight line toward LMC~X--3. \\citet{wang05} have reported the detection of the hot gas associated with our Galaxy, based on X-ray and far-UV absorption line spectra from \\chandra\\ and \\fuse\\ observations. Here we present the {\\sl Suzaku} CCD emission spectra of the X-ray background in two fields adjacent to the LMC~X--3 sight line. This unique combination of the X-ray and far-UV spectral data toward essentially the same part of the sky further allows us to constrain the spatial, thermal, and chemical properties of the hot gas. This paper is organized as follows. We present the {\\sl Suzaku} observations and the data calibration, as well as a brief description of the existing {\\sl Chandra} and {\\sl FUSE} data in \\S~\\ref{sec:obs}. In \\S~\\ref{sec:results}, we first analyze the X-ray emission ({\\S~\\ref{sec:emission}) and absorption ({\\S~\\ref{sec:abs}) data separately, and then build a slab-like hot gas model to jointly analyze these X-ray data (\\S~\\ref{sec:model}). In \\S~\\ref{sec:fuse}, we compare the model predicted \\ovi\\ absorptions with the \\fuse\\ detections, and then use the observed \\ovi\\ absorption line to further constrain our model. We discuss the implications of our results in \\S~\\ref{sec:dis} and summarize our results and conclusions in \\S~\\ref{sec:sum}. Throughout the paper, we assume the hot emitting/absorbing gas to be optically thin (see \\S~\\ref{sec:scattering} for further discussion) and in the CIE state. We adopt the solar abundances from \\citet{and89} and quote parameter errors at 90\\% confidence levels for a single varying parameter unless otherwise noted. We also refer the hot gas on scales of several kpc as the Galactic disk, in contrast to the Galactic halo on scales of $>10$ kpc (see \\S~\\ref{sec:location} for further discussion). Our spectral analysis uses the software package XSPEC (version 11.3.2). ", "conclusions": "\\label{sec:dis} We have presented the high quality {\\sl Suzaku} emission observations of the diffuse hot gas toward two off-fields of the LMC~X--3 sight line. In particular, the \\ovii\\ and \\oviii\\ emission lines are clearly resolved. Modeling these emission spectra yields a gas temperature that is about two times higher than that inferred from the high resolution absorption data. We find that our non-isothermal thick Galactic gaseous disk model can account for this discrepancy as well as the far-UV \\ovi\\ absorption data. A joint fit to these data indicates that both the X-ray absorption and emission and the far-UV absorption are consistent with being produced from hot gas in a region of several kpc around the Galactic plane. In the following, we first discuss how our results are potentially affected by caveats in our data analysis, and then discuss the implications of our results on the origin and cooling of the \\ovii- and \\ovi-bearing gases. \\subsection{Uncertainty of the LHB and the SWCX} \\label{sec:LHB} The biggest uncertainty in modeling the SXB emission is the estimate of the contributions from the LHB and SWCX. Our understanding for both components is still very poor because it is very hard to decompose their contributions. The SWCX emission could in principle be estimated by tracing the solar wind flux and the neutral atoms in the Earth's atmosphere and in the local ISM (e.g., Lallement \\etal 2004). However, such an estimation is model dependent. For instance, for the line of sight with Galactic coordinates $(l, b)=(278.65^\\circ, -45.30^\\circ)$ observed on 2003 March 3, the \\ovii\\ emission is estimated to be 0.83 and 3.7 LU by \\citet{kou07} and by \\citet{hen08}, respectively. Recent comparisons of the observed X-ray emission with models for the SWCX suggest that the foreground emission in the shadowing experiments can be mainly attributed to the SWCX and that the LHB emission is negligible (Koutroumpa \\etal 2007; Rocks \\etal 2007; but also see Shelton 2008). The total \\ovii\\ intensity of the SXB toward a high Galactic latitude sight line ($l, b = 272^\\circ, -58^\\circ$) is 2.7(2.0, 3.4) LU (1$\\sigma$ range), obtained from a recent {\\sl Suzaku} observation \\citep{mit06}. If emission from both the LHB and the SWCX is isotropic and independent of Galactic latitude, this value can be regarded as the upper limit of the combined contribution of these two components. Our results are not qualitatively affected by above discussed uncertainties. To be more quantitative in assessing the effect, we allow EM of the LHB plus SWCX component to vary from 0 to 0.0113 ${\\rm cm^{-6}~pc}$ (\\S~\\ref{sec:emission}), representing two extreme cases. The low boundary corresponds {\\it zero} emission of the LHB and the SWCX components, whereas the high boundary corresponds to the 3$\\sigma$ upper limit of the observed \\ovii\\ intensity (i.e., 4.8 LU) toward the high Galactic latitude sight line. We then re-perform our spectral analysis described in \\S\\S~\\ref{sec:emission} and \\ref{sec:fuse}. We find that new constrained gas temperature in modeling the emission data alone under the isothermal assumption is still about two times higher than that constrained from the absorption data (refer to the first and the fifth rows in Table~\\ref{tab:results}), which still necessitates the non-isothermal model developed in \\S~\\ref{sec:model}. The new results constrained in the joint analysis are also consistent with the old values, except for with a slightly broader uncertain range (refer to the forth and the sixth rows in Table~\\ref{tab:results}). \\subsection{Uncertainty of the extragalactic power-law component} \\label{sec:cxb} In modeling the emission spectra, we used a simple PL function to approximate the extragalactic contribution (\\S~\\ref{sec:emission}). Recent {\\sl Chandra} and {\\sl XMM-Newton} deep surveys for the cosmic X-ray background have resolved $\\gsim80\\%$ of the background emission at 1-8 keV into extragalactic discrete sources (e.g., Hickox \\& Makevitch 2006). Optically bright sources tend to be spectrally hard while faint sources tend to be spectrally soft (e.g., Mushotzky \\etal 2000). To examine the goodness of our approximation, we replace the simple PL with a broken PL and fix the break energy at 1 keV to model the extragalactic component in our joint analysis described in \\S~\\ref{sec:model}. We obtain the photon indices as 1.7(1.3, 2.0) and 1.4(1.3, 1.5) below and above 1 keV, respectively, and find that the parameters of the Galactic hot gas are barely affected (the seventh row of Table~\\ref{tab:results}). \\subsection{Effect of resonant scattering of the \\ovii\\ line emission} \\label{sec:scattering} In this work, we have assumed the emitting/absorbing gas to be optically thin, i.e., we have neglected the resonant scattering effect of the hot gas. To qualitatively assess the validity of our assumption, we assume the hot gas density to be uniform for easy of analysis. The observed \\ovii\\ emission line consists of unresolved triplet transitions, resonance, forbidden, and intercombination lines at 21.60 \\AA, 22.10 \\AA, and 21.80 \\AA, respectively. Since the oscillation strength is essentially zero for the latter two transitions, we consider the ``scattering'' only for the resonant line. The absorbed resonant photons should be re-emitted isotropically; half of these re-emitted photons are expected to favor the observing direction. Taking all these into account, we obtain the observed line intensity as \\begin{equation} I = (1 - f) I_0 + 0.5fI_0 \\left( 1 + \\frac { 1 - e^{-\\tau} }{\\tau} \\right ), \\end{equation} where $\\tau$ is the (max) absorption optical depth at center wavelength of the resonant transition, and $I_0$ is the ``intrinsic'' (without the scattering) line intensity. Here, $f$ is the resonant transition fraction of the \\ovii\\ triple, which is a function of gas temperature. Taking the (max) temperature at the Galactic plane $T_0\\sim3\\times10^6$ K (Table~\\ref{tab:results}), which favors more to the resonant transition and gives $f=0.65$, we get $I \\gsim 0.675I_0$ for any value of $\\tau$. This indicates that the observed intensity should not be different from the intrinsic value by a factor larger than 1.5. In reality, the ``background'' gas cloud also scatters the photons emitted by the ``foreground'' cloud to the observing direction. If this effect is further considered, the difference should largely disappear. Therefore we conclude that the resonance scattering should not significantly affect our results, although a more detailed modeling needs to be performed to account for a more realistic geometry of the hot gas distribution. \\subsection{Origin of the \\ovii-bearing gas in general} \\label{sec:location} Our analysis indicates that the gas responsible for the observed X-ray absorption/emission along the LMC~X--3 sight line has a pathlength about a few kpc and is consistent with the Galactic disk in origin. This result has important implications for understanding the structure and origin of hot gas around our Galaxy. As described in \\S~\\ref{sec:intro}, high ionization X-ray absorption lines with zero velocity shift have been observed along many extragalactic sight lines, and the \\ovii\\ emission lines are also detected at various high Galactic latitudes. Several scenarios have been proposed for the origin of the emitting/absorbing gas, including the intergalactic medium (IGM) of the Local Group, the large-scale Galactic halo, and the thick Galactic gaseous disk (see Yao \\& Wang [2007] for a review). These scenarios cannot be distinguished kinematically because of the limited spectral resolution of current X-ray instruments. So currently the most direct information about the location of the gas comes from differential analysis of the X-ray absorption lines toward sources at different distances. \\citet{wang05} find that the \\ovii\\ absorption along the sight line toward LMC~X--3, at a distance of $\\sim50$ kpc, is comparable to those observed toward AGN sight lines. Therefore, if the LMC~X--3 sight line is representative, one can then conclude that the bulk of the X-ray absorption around the Galaxy is within $\\sim50$ kpc. A similar conclusion has been reached based on the estimate of the total baryonic matter contained in the \\ovii-bearing gas, and on the angular distribution of the \\ovii\\ absorption \\citep{fang06, bre07}. A comparison of a Galactic sight line (4U~1957+11) with two extragalactic sight lines toward LMC~X--3 and Mrk~421 further indicates that there is no significant \\ovii\\ absorption in the extended Galactic halo beyond a few kpc \\citep{yao08}, which is consistent with the conclusion drawn in this work. The \\ovii\\ emission from the hot gas beyond LMC~X--3 is not expected to be important either. \\citet{yao08} obtained a 95\\% upper limit to the \\ovii\\ absorption beyond LMC~X--3 as $N{\\rm _{OVII}<3.7\\times10^{15}~cm^{-2}}$. Assuming this much of \\ovii\\ uniformly distributed in a region with an extension scale of $l$, we calculate its emission as $I_{\\rm OVII}<0.025/(A_{\\rm O0.1}l_{\\rm 100kpc})$ LU for the entire temperature range from $10^5$ to $10^7$ K, where $A_{\\rm O0.1}$ is the oxygen abundance in unit of 10\\% solar value and $l_{\\rm 100kpc}$ is in unit of 100 kpc. In contrast, the inferred the \\ovii\\ intensity of the Galactic disk diffuse gas is 5.0 LU (\\S~\\ref{sec:emission}; Table~\\ref{tab:line}). We have shown that the X-ray emission and absorption as well as the far-UV \\ovi\\ absorption along the sight line toward \\xs\\ can be explained by the presence of a thick Galactic hot gaseous disk. A similar explanation holds for the sight line toward Mrk~421, for which another joint X-ray absorption/emission and UV-absorption has been carried out (the X-ray emission is, however, mostly based on the RASS broad-band measurements; Yao \\& Wang 2007). The characteristic scale height of the disk is also consistent with the \\neix\\ column density distribution in the Galaxy \\citep{yao05}. The gas in the disk is shown to have a normal metal abundances \\citep{wang05, yao06} and is clearly due to the heating by mechanical energy feedback from stars in the Galaxy (e.g., Ferri\\`ere [1998]). Such an interpretation is also supported by the X-ray emission observations of nearby disk galaxies like our own. The diffuse X-ray emission has been routinely observed in many late-type normal star-forming spirals and is confined to a region around galaxies with vertical extent no more than a few kpc (e.g., Wang \\etal 2001, 2003; T\\\"ullmann \\etal 2006a, 2006b). Both the intensity and the spatial extent appear to be scaled with galactic star formation rate (e.g., T\\\"ullmann \\etal 2006b). Therefore, we conclude that the X-ray emitting/absorbing and \\ovi-absorbing gas, as studied in this paper, arises primarily from the stellar feedback in the Galactic disk. \\subsection{Location of the \\ovi-bearing gas} \\label{sec:o6location} Before discussing the \\ovi-bearing gas location, let us look at the \\ovi\\ emission properties first. In \\S~\\ref{sec:fuse} we have shown that the observed \\ovi\\ absorption can be predicted from the X-ray-data-constrained non-isothermal gas model and have then used the \\ovi\\ $\\lambda$1031.96 absorption line to further constrain our spectral model. With the fitted model parameters, we can estimate the intrinsic \\ovi\\ emission intensity of the diffuse hot gas toward LMC~X--3, and then compare it with observations. The best-fit model gives 186 LU in total for the \\ovi\\ doublet $\\lambda\\lambda$1032 and 1038. Taking all the 90\\% boundaries that favor more \\ovi\\ emission, we obtain a firm upper limit of 1612 LU. There is no direct measurement of the \\ovi\\ emission within 1$^\\circ$ of the LMC~X--3 sight line. Within $\\sim5^\\circ$, some observations show measurable \\ovi~$\\lambda1032$ emission as $2.1-5.7 \\times10^3$ LU, while other observations only yield upper limit even with longer exposures \\citep{ott06, dix06, dix08}. Since the \\ovi\\ emission could vary by a factor of $>2.3$ at an angular scale as small as $25'$ \\citep{dix06}, it is nearly impossible to reliably estimate the \\ovi\\ emission toward LMC~X--3 sight line based on the few existing nearby samples. However, if there were a measurable \\ovi\\ emission toward the LMC~X--3 sight line (but see below), it must be in order of several thousand LU, as typically observed at high galactic latitudes, which is significantly larger than that predicted in the \\ovii-bearing gas and therefore unlikely arises from the diffuse hot gas. Where is the \\ovi-bearing gas then? To answer this question, it is important to compare the spatial distributions of the \\ovi\\ absorption and emission. In this work, we find that most of the \\ovi\\ absorption can arise from the diffuse gas along a vertical scale of $\\sim$3 kpc around the Galactic plane. Figure~\\ref{fig:OVI_dis} shows the density and column density distributions of \\ovi\\ and \\ovii\\ as a function of the vertical distance, predicted from our constrained non-isothermal disk model. While compared to \\ovii, \\ovi\\ is distributed in a relative narrow region, which is mainly due to the sensitive dependence of \\ovi\\ population on temperature (Fig.~\\ref{fig:absVSemi}), about 80\\% of the \\ovi\\ column still spreads over as large as 1 (from 2.7 to 3.7) kpc. Observationally, \\ovi\\ absorption has been detected toward essentially all sight lines, and the characteristic scale height of the \\ovi\\ column densities is similar to that constrained in this work (e.g., Savage \\etal 2003). It has been suggested that \\ovi\\ could largely arise at interfaces between hot and cool interstellar media. However, for Mrk~421 sight line, \\citet{sav05} counted the number of cool gas velocity components and found that the interfaces can only account for at most 50\\% of \\ovi\\ absorption, leaving a bulk of the \\ovi\\ to other origins. In contrast, the \\ovi\\ emission is measured toward only $\\sim30\\%$ of the surveyed sky \\citep{ott06, dix08}. These results clearly indicate that the ``commonly'' observed \\ovi-absorbing gas in kpc scale is not primarily responsible for the detected \\ovi\\ emission. We propose that the low temperature regime of the kpc-scale diffuse gas could contain a significant or even dominant fraction of the observed \\ovi\\ absorption, as indicated in this work, and that the conductive interfaces between hot and cool gases, though could be another reservoir of \\ovi\\ ion, is mainly responsible for the observed \\ovi\\ emission. This picture not only explains the spatial distribution of the \\ovi\\ absorption, but also naturally explains the sight line to sight line \\ovi\\ variation in both absorption and emission. For LMC~X--3 sight line in particular, since all the observed \\ovi\\ can be attributed to the diffuse gas, the \\ovi\\ emission is not expected to be strong. \\begin{figure} \\plotone{f7.eps} \\caption{Spatial distribution of density (a) and column density (b) of \\ovi\\ and \\ovii, predicted with our constrained non-isothermal disk model. \\label{fig:OVI_dis} } \\end{figure} \\subsection{Radiative energy loss rate of the hot gas} \\label{sec:lossrate} If hot gas toward the LMC~X--3 sight line represents reasonable well the hot ISM of the Galactic disk as a whole, we can then use our non-isothermal gas model to estimate the total radiative energy loss rate of the gas, and compare it to the expected energy input from SN explosions. It has long been observed that in the low $L{\\rm_x}/L{\\rm_B}$ ($L_{\\rm B}$ is the blue luminosity) early-type galaxies, X-ray radiative energy loss rate of the diffuse hot gas is far less than the expected SN energy input (e.g., Canizares \\etal 1987; Brown \\& Bregman 2000). However, all the previous studies were based on imaging observations in which decomposing different emission components is a complicated task (e.g., Li \\etal 2007) and could make the inferred properties of the ``true'' diffuse hot gas very uncertain. The LMC~X--3 sight line is so far the only sight line along which the high resolution X-ray emission/absorption and far-UV absorption measurements with a minimal confusion are available, which allow us to tightly constrain the thermal and chemical properties as well as the global spatial distribution of the Galactic diffuse hot gas, as presented in this work. It is thus worthwhile reexamining the radiative power of the Galactic hot gas based on these characterizations. The energy loss rate of the hot gas can be expressed as \\begin{equation} \\label{equ:lossrate} \\frac{d^2U}{dsdt} = 2\\times\\int_0^\\infty 1.2n^2\\Lambda(T)~dz, \\end{equation} where the factor 2 accounts for the two sides of the Galactic disk. With our model (Eq.~\\ref{equ:expHT}), we can rewrite the Eq.~\\ref{equ:lossrate} as \\begin{equation} \\label{equ:rate} \\frac{d^2U}{dtds} = 2\\times\\int^{T_0}_{T_{min}} 1.2n_0^2 \\left(\\frac{T}{T_0}\\right)^{2\\gamma-1}\\Lambda(T) (-h_T)~d\\left(\\frac{T}{T_0}\\right). \\end{equation} Assuming the solar abundances, plugging in the best-fit $n_0$, $T_0$, $\\gamma$, and $h_T$ values (\\S~\\ref{sec:model}), and taking the emissivity $\\Lambda(T)$ from the atomic database ATOMDB\\footnote{http://cxc.harvard.edu/atomdb}, we obtain the local surface emissivity of the hot gas as 3.1 and 16.1 $\\times10^{36}~{\\rm ergs~s^{-1}~kpc^{-2}}$ in 0.1-10 and 0.01-10 keV bands, respectively. Including the additional \\ovi\\ emission of $\\sim$3000 LU from the interface component, the total radiative energy loss rate is $\\sim3\\times10^{37}~{\\rm ergs~s^{-1}~kpc^{-2}}$. The radiative cooling of the Galactic disk hot gas only accounts for less than 10 per cent of the expected energy input of the stellar feedback. At Sun's galactocentric radius, the SN rate is 19 and 2.6 ${\\rm Myr^{-1}~kpc^{-2}}$ for type II and type Ia, respectively \\citep{fer98}. If on average each type II SN progenitor releases $2\\times10^{50}$ ergs of energy before it explodes and each SN explosion releases $10^{51}$ ergs \\citep{fer98, lei92}, the total energy input is then $8\\times10^{38}~{\\rm ergs~s^{-1}~kpc^{-2}}$, which is about 20 times higher than our estimated gas cooling rate. The ``missing'' energy could be either emitted in other wavelength bands (e.g., infrared) or have been consumed in driving other Galactic activities (e.g., galaxy-size out flows). Obviously, such an estimation is model dependent. \\citet{she07} recently concluded that $\\sim70\\%$ of the SN energy input could be radiated away by the Galactic hot gas. In their model, they required the observed \\ovi\\ emission to be co-spatial with the \\ovi\\ absorption and assumed both the \\ovi-bearing and the hotter \\ovii-bearing gas to be isobaric. This requirement results in that most of the \\ovi-bearing gas is confined within a region of $<100$ pc in size, which favors to produce about ten times more thermal emission in the energy range of 0.01-0.1 keV than our model. In contrast, we constrain our model with X-ray \\ovii\\ and \\oviii\\ emission and absorption, and find that the most of the \\ovi\\ absorption could arise from the kpc-scale diffuse gas. In our model, because of large extent of the diffuse gas (therefore a low density), its emitting power is at most as much as that of the observed/expected \\ovi\\ emission lines, which presumably arise from the interfaces between the hot and cool interstellar media (\\S~\\ref{sec:o6location}). The scale size of the \\ovii- and \\ovi-bearing gas derived in this work and utilized in our estimation, is consistent with that obtained in the recent \\ovi\\ absorption surveys \\citep{sav03, bow08}. \\subsection{Evidence for oxygen and iron depletion?} \\label{sec:depletion} In our analysis, we find that, while the relative abundance of Fe/O is about the solar value, the Ne/O is about 2 times higher (Table~\\ref{tab:results}). This overabundance of neon, if true, is consistent with the interpretation of about 50 per cent of oxygen and iron being depleted into dust grains. However, because this overabundance is mainly derived from the emission data in which decomposing various components is still very uncertain, the depletion interpretation therefore may not be unique. For instance, \\citet{hen08} suggest that the SWCX could also produce apparent non-solar abundance ratio. Furthermore, our understanding of the chemical abundances in the solar system is still poor. For example, the oxygen value has recently been revised downward by $\\sim35\\%$ \\citep{asp05}, but this revision is still under debate \\citep{ant06}. The abundance pattern in the ISM is also suggested to be slightly different from that in the solar system \\citep{wil00}. So it is more useful to list the number density ratio of Ne/O derived in this work, which is 0.25(0.19, 0.33). We prefer to defer the interpretation of the apparent overabundance of neon until a better understanding of both the SWCX and the solar chemical abundances is reached." }, "0808/0808.1980_arXiv.txt": { "abstract": "We present a new method for constructing equilibrium phase models for stellar systems, which we call the iterative method. It relies on constrained, or guided evolution, so that the equilibrium solution has a number of desired parameters and/or constraints. This method is very powerful, to a large extent due to its simplicity. It can be used for mass distributions with an arbitrary geometry and a large variety of kinematical constraints. We present several examples illustrating it. Applications of this method include the creation of initial conditions for $N$-body simulations and the modelling of galaxies from their photometric and kinematic observations. ", "introduction": "In astronomy there are at least two problems where equilibrium phase models of stellar systems need to be constructed. The first one is the construction of phase models for real galaxies from observational data, i.e. the modelling of observational data. The second problem is the construction of initial conditions for $N$-body simulations of stellar systems. It is obvious that these two problems are tightly connected, and yet they have, so far, been solved by different methods. The Schwarzschild method \\citep{S79} and its modifications is often used for modelling of observational data \\citep[e.g][]{H00, vdB06, T07, vdB08, deL08}, but has almost never been used so far to produce initial conditions for simulations. For $N$-body initial conditions, a wide variety of methods has been used, based on the Jeans theorem \\citep[e.g. ][]{Zangthesis, AS86, KD95, WD05, M07}, or on Jeans' equations \\citep[e.g. ][]{H93}. In the case of multi-component systems, e.g. disc galaxies with a bulge and a halo, the components are built separately and then either simply superposed, or the potential of the one is adiabatically grown in the other \\citep[e.g. ][]{Barnes88, M07, A07} before superposition. For real galaxies the phase space density is generally unknown, but we do have some information about it. For example, we know more or less accurately a distribution of mass for the visible components (notwithstanding uncertainties due to the mass to light ratio) and we often have some constraints on the velocity distribution. It is also reasonable to assume that the galaxy is in an equilibrium state. So in general, the problem of constructing a model in phase space is equivalent to constructing an equilibrium phase model with a given mass distribution and, in many cases, given kinematic constraints. In the case of modelling observational data (first of the two above mentioned problems) the kinematic parameters are the observed velocities integrated along the line of sight. In the case of initial conditions for $N$-body simulations, a wide variety of kinematic parameters is possible, depending on the problem the simulation addresses. We have developed a new method for constructing equilibrium phase models with a given mass distribution and with given kinematic parameters, which we call the iterative method. It can be applied to systems with arbitrary geometry, so that the requested mass distribution can be arbitrary. The idea and a first implementation was presented in \\citet{RS06}. In \\citet{RO08} we improved it, and applied it to construct an $N$-body model of the stellar disk of our Galaxy for two realistic mass models of the Milky Way. Here we present a final version of this method, fully allowing kinematical constraints. In the previous articles we had concentrated on constructing equilibrium phase models with a given mass distribution, so that kinematic parameters were either not considered or only in terms of auxiliary parameters, such as the total angular momentum \\citep{RS06, RO08}. This, however, limited the applicability of our method, both for initial conditions and for modelling real galaxies. Indeed, initial kinematics play a crucial role in determining the evolution of $N$-body systems, while observational constraints more often than not include kinematics. In this paper we give equal attention to the mass distribution and kinematical constraints, so that the iterative method can now be used for a number of interesting applications. In principle, in our method, both the kinematic constraints and the mass distribution can be arbitrary. But the part of our algorithm that concerns the kinematic constraints is not universal, contrary to the part that handles the mass distribution, but is tailored to the specific constraint. Here we consider several types of constraints. Once these are understood, it is rather easy to extend the algorithm for every new type of kinematic parameter (see below). The power of the iterative method stems from its simplicity. The iterative method is based on a simple and, in a way, obvious idea, which is implemented in a simple algorithm. The purpose of this article is to fully describe this method. We first introduce the basic concept in Section~\\ref{s_imethod}, where we also explain the different modules of the algorithm and the way they should be applied. In section~\\ref{s_models} we illustrate the use of the method with three examples, namely a triaxial system, a multi-component model of a disk galaxy (including live disk, bulge and halo components) and a disk constructed with given line-of-sight kinematic. We briefly conclude in section~\\ref{s_conc}. ", "conclusions": "\\label{s_conc} We presented a new method for constructing equilibrium phase models for stellar systems --- the iterative method. The aim of this method is to construct equilibrium $N$-body models with given parameters, or constraints. More specifically, these are a given mass distribution and, if desired, given kinematic properties, parameters, or constraints. Our method is straightforward both conceptually and in its implementation. We believe that it is this simplicity that makes this method so powerful. It simply relies on a constrained, or guided evolution. We let the system reach equilibrium via a dynamical evolution in a number of successive steps. In between two such steps we make sure that the parameters are set to their desired value and/or that the constraints are fulfilled. This means that the evolution is guided towards an equilibrium with the desired parameters and/or constraints. Setting a mass distribution is of course obligatory, but kinematical constraints are not. If we wish to include them, we have the choice of a large number of possibilities, such as setting the radial profile(s) of one, or more moments of the velocity distribution. In this article we described only a few types of kinematic constraints: the profile of radial velocity dispersion, the profile of velocity anisotropy, a condition of velocity isotropy and line-of-sight kinematics. Procedures for further types of kinematic constraints can be easily found following similar techniques. Furthermore, our implementation of the iterative method can be directly applied to systems with arbitrary geometry, i.e. the given mass distribution can be arbitrary and need not have any symmetries. Thus our method can be used in many different applications. We used our iterative method to construct several models. The first one is a triaxial system. The second one is a multi-component model of a disk galaxy consisting of a stellar disk with a given radial velocity dispersion profile, a non-spherical bulge and a halo with a given anisotropy profile. We also constructed two disk models with given line-of-sight kinematic. Using self-consistent $N$-body simulations, we made sure that the models we constructed are indeed very close to equilibrium (see figs.~\\ref{fig_3d},~\\ref{fig_disk},~\\ref{fig_bulge}, ~\\ref{fig_halo},~\\ref{fig_svlos_disk} and~\\ref{fig_mvlos_disk}). The iterative method has a number of further applications. It can of course be used for constructing equilibrium initial conditions for $N$-body modelling of stellar systems. For instance, the iterative method allows one to investigate bar formation in galaxies with a halo having different kinematics. Also the iterative method can be applied for constructing phase models of real galaxies. For example, we can model observational data by constructing phase models with given line-of-sight kinematics, as shown in section~\\ref{s_losdisks}. This paves the way for studies of e.g. the distribution of dark matter in ellipticals, or obtaining phase space models of observed disk galaxies. A further interesting application is the study of the properties of several equilibrium models for a given mass distribution, as for example triaxial systems. The software necessary for the implementation of this method should be thought of in a very modular way, e.g. with different units for the various kinematical constraints, and is very straightforward to write. Nevertheless, we will make our own software publicly available as soon as this paper is accepted, at the address http://www.astro.spbu.ru/staff/seger/soft/. This package will contain also step-by-step examples for constructing models by using the iterative method, including the models described in this article. Our software uses the $N$-body code gyrfalcON \\citep{D00, D02} and the NEMO package (http://astro.udm.edu/nemo; \\citealt{T95})." }, "0808/0808.1949_arXiv.txt": { "abstract": "Stellar jets are normally constituted by chains of knots with some periodicity in their spatial distribution, corresponding to a variability of order of several years in the ejection from the protostar/disk system. A widely accepted theory for the presence of knots is related to the generation of internal working surfaces due to variations in the jet ejection velocity. In this paper we study the effect of variations in the inner disk-wind radius on the jet ejection velocity. We show that a small variation in the inner disk-wind radius produce a variation in the jet velocity large enough to generate the observed knots. We also show that the variation in the inner radius may be related to a variation of the stellar magnetic field. ", "introduction": "Stellar jets are observed in the form of chains of emitting nebulae called Herbig-Haro (HH) objects. HH objects are generally thought to be shock-heated density condensations traveling along the outflows formed by star-disk systems during the process of star formation \\citep[e.g.][]{rei01}. Since their discovery several scenarios have been suggested for the origin of \\emph{steady} outflows from young stellar objects. In the stellar wind model, material is accelerated by thermal pressure gradients \\citep[e.g.][]{can80}. Magnetohydrodynamics (MHD) models rely on the magnetocentrifugal launching mechanism \\citep{bla82}. For the X-wind scenario the jet is magnetically driven from the so-called ``X-annulus'' where the young star's magnetosphere interacts with the disk \\citep{shu00}. In the disk wind scenario the jet is launched from an extended region of the disk surface \\citep[e.g.][]{fer97}. The analytical models mentioned above have focused mainly on the steady-state aspect of the ejection phenomena. \\emph{Unsteady} periodic ejections with timescales of the order of several rotation periods of the inner disk radius have been obtained by numerical simulations \\citep[e.g.][]{ouy97,goo99,mat02}. Observations indicate a significantly longer timescale is associated with the appearance of knots in stellar outflows. Nearly all observed jets present small scale knots up to 0.1 pc from the central source, with a spacing between the knots corresponding to a timescale of $\\approx$ 1-20 yr (e.g. HH30 - \\citealt{bur96}; HH111 - \\citealt{rei92}; RW-Aur - \\citealt{lop03}). More fragmented knots are observed on a typical timescale of $\\sim 10^{2-3}$ yr at larger distances from the source. While the long term variation of jets can be explained by variations in the accretion rates (e.g. during FU-Orionis phases), the possible origin of small scale knots is still unclear. A first possibility, suggested by similarities with extra-galactic jets, is that the knots may be formed by hydrodynamics Kelvin-Helmholtz instabilities \\citep{mic98}, MHD Kelvin-Helmholtz reflective pinch mode instabilities \\citep[]{cer99}, or by current-driven instabilities \\citep[]{fra00}. Numerical simulations by these authors have shown that the shocks generated by plasma instabilities are weaker than those seen in observations once radiative cooling is taken into account. Moreover, some jets \\citep[e.g. HH212 -][]{zin98} show a remarkable symmetry on both sides of the central star-disk system. This may be difficult to explain assuming that the instabilities are triggered by small scale perturbations. Additionally, and more importantly, optical images in many cases show that the compact knots have a bow-shock structure (e.g. HH111 - \\citealt{rei92}) that is difficult to obtain supposing that the knots are formed by instabilities. A widely accepted theory for the presence of knots in stellar jets is related to the generation of internal working surfaces due to supersonic variations of the ejection velocity at the base of the jet \\citep{rag90}. Indeed, numerical simulations, using a velocity variations of about 10-20\\% of the average velocity, have been able to reproduce the morphology and the emission property of the knots in HH objects \\citep[e.g.][]{esq07}. In this paper we show that velocity variations which lead to the creation of knots similar to the observed ones may be generated by variations of the inner disk-wind radius. Furthermore, we suggest that these variations may be related to changes in the stellar magnetic field. This paper is organized as follows. In Section \\ref{model} we determine the effect of a variation in the inner disk radius on the jet velocity. In Section \\ref{variable} we suggest that the variation in the inner radius could be connected to a change of the stellar magnetic field. In Section \\ref{discussion} we discuss the approximation used and the limitations of our model. Conclusions are given in Section \\ref{conclusion}. ", "conclusions": "\\label{conclusion} In this paper we explored the effect of a stellar magnetic field variation on the ejection velocity of protostellar jets. While large scale knots may successfully be explained by a large increase in the accretion rate, we showed that small scale knots may be produced by changes in the inner radius. We showed that a stellar magnetic field variation, if present, represents a natural candidate to produce stellar knots similar to the observed. In fact, a stellar magnetic field variation produces a change in the inner radius and therefore in the jet velocity. We also estimated that a stellar magnetic field variation $\\lesssim$ 50\\% produces a variation ($\\sim$ 20\\%) in the average velocity large enough to produce knots similar to the observed. The stellar magnetic field periodic or quasi-periodic variation remains an hypothesis of our scenario, but future large scale temporal observations of magnetic field in young stars, and progress in dynamo theory, will both help to provide an answer to this problem." }, "0808/0808.1208_arXiv.txt": { "abstract": "Previous studies of phase mixing of ion cyclotron (IC), Alfv\\'enic, waves in the collisionless regime have established the generation of parallel electric field and hence acceleration of electrons in the regions of transverse density inhomogeneity. However, outstanding issues were left open. Here we use 2.5D, relativistic, fully electromagnetic PIC (Particle-In-Cell) code and an analytic MHD (Magnetohydrodynamic) formulation, to establish the following points: (i) Using the generalised Ohm's law we find that the parallel electric field is supported mostly by the electron pressure tensor, with a smaller contribution from the electron inertia term. (ii) The generated parallel electric field and the fraction of accelerated electrons are independent of the IC wave frequency remaining at a level of six orders of magnitude larger than the Dreicer value and approximately 20 \\% respectively. The generated parallel electric field and the fraction of accelerated electrons increase with the increase of IC wave amplitude. The generated parallel electric field seems to be independent of plasma beta, while the fraction of accelerated electrons strongly increases with the decrease of plasma beta (for plasma beta of 0.0001 the fraction of accelerated electrons can be as large as 47 \\%). (iii) In the collisionless regime IC wave dissipation length (that is defined as the distance over which the wave damps) variation with the driving frequency shows a deviation from the analytical MHD result, which we attribute to a possible frequency dependence of the effective resistivity. (iv) Effective anomalous resistivity, inferred from our numerical simulations, is at least four orders of magnitude larger than the classical Spitzer value. ", "introduction": "Phase mixing is a mechanism of enhanced dissipation of Alfv\\'en waves due to inhomogeneity of Alfv\\'en speed in a direction transverse to a local magnetic field. This mechanism originally was studied in the fusion and laboratory plasma context by a number of authors \\cite{1972PhFl...15.1673U,1973ZPhy..261..203T,1973ZPhy..261..217G,1974PhRvL..32..454H, 1974PhFl...17.1399C,1975JPlPh..13...87T} and subsequently applied to the solar corona \\cite{1983A&A...117..220H}. Most of the large amount of work done in the field of phase mixing was in the resistive MHD (Magnetohydrodynamic) regime. Recently, a few works looked at the same mechanism in the collisionless regime in the context of Earth magnetosphere \\cite{1999JGR...10422649G, 2004AnGeo..22.2081G} and solar corona \\cite{2005A&A...435.1105T,2005NJPh....7...79T}. The main findings of these works include the generation of electric field that is parallel to the ambient magnetic field in the regions of transverse density inhomogeneity, as well as associated electron acceleration. It should be mentioned that these studies considered circularly polarised ion cyclotron (IC) waves which in the low frequency regime become Alfv\\'en waves. We use terms Alfv\\'en or IC interchangeably, but reader should bear in mind we always refer to {\\it waves with frequencies $< \\omega_{ci}$} (with $\\omega_{ci}$ being ion cyclotron frequency). The exact mechanism of generation of the parallel electric field has stimulated a debate \\cite{2006A&A...449..449M,2007NJPh....9..262T}, and even MHD regime option was explored \\cite{2006A&A...455.1073T}. Continuing this investigation, here we apply technique used in the collisionless reconnection \\cite{pritchett01,th08}. Namely, in section 3.1 we use generalised Ohm's law to find out which term generates the parallel electric field. Solar flare observations \\cite{2005SSRv..121..141F} trigger one's interest in how effectively plasma particles are accelerated. Hence, in section 3.2 we look into how the generated parallel electric field and the fraction of accelerated electrons depend on model parameters such as IC wave frequency, amplitude, and plasma beta. Ref. \\cite{1983A&A...117..220H} provides a simple analytical expression how Alfv\\'en wave amplitude should decay in space due to the phase mixing. Despite the fact that their formula is derived in the resistive MHD regime, we still apply it to our collisionless, kinetic simulation and see what does the comparison yield (Section 3.3). This is done in the light of previous results of \\citet{2005A&A...435.1105T} who established that in the collisionless, kinetic regime Alfv\\'en wave amplitude in the density gradient regions decays with distance (from where it is driven) according to collisional MHD formula of \\citet{1983A&A...117..220H}. Here we stretch the MHD-kinetic analogy further to test $\\omega_d^2$ dependence under the exponent. In Sect 3.4 we estimate the effective \"resistivity\" (again the spirit of MHD-kinetic analogy). The quotation marks are needed to signify that PIC (Particle-In-Cell) simulation code is collisionless and hence no resistive effects exist as such. However, scattering of particles by magnetic fields plays effective role of collisions. ", "conclusions": "Let us summarise the above findings: We used the generalised Ohm's law and found that the parallel electric field, which is generated by propagation of IC (Alfv\\'enic) wave in a transversely inhomogeneous plasma, is supported mostly by the electron pressure tensor, with a smaller contribution from the electron inertia term. Surprisingly, this result resembles closely to the previous results on collisionless reconnection both in tearing unstable Harris current sheet \\cite{hesse99,birn01,pritchett01} and stressed X-point collapse \\cite{th07,th08}. However, in the latter two cases, the generated electric field is in the plane perpendicular to the magnetic field. Thus, a universal importance of the electron pressure tensor in relation to supporting the electric fields in collisionless plasmas should be noted. We explored physical parameter space of the problem with regards to the efficiency of generation of parallel electric field and acceleration of electrons. We found that the generated parallel electric field and the fraction of accelerated electrons are independent of the IC wave frequency staying at a level that is $10^6$ times larger than the Dreicer value and approximately 20 \\% respectively. The generated parallel electric field and the fraction of accelerated electrons increase with the increase of IC wave amplitude. The generated parallel electric field seems to be independent of plasma-$\\beta$. However, the fraction of accelerated electrons strongly increases with the decrease of plasma-$\\beta$, e.g. for plasma $\\beta=0.0001$ the fraction of accelerated electrons can be as large as 47 \\%. Previously it was established that in the collisionless, kinetic regime phase-mixed Alfv\\'en (IC) wave amplitude damps with distance of propagation according to $\\propto \\exp[-(x/L_d)^3]$ \\cite{2005A&A...435.1105T}, which resembles closely to collisional MHD result of \\citet{1983A&A...117..220H}. We tried to stretch this analogy further by investigating how the dissipation length $L_d$ scales with the IC driving frequency. We found that the scaling is different from the MHD result. We have shown that this discrepancy can be attributed to the frequency dependence of the effective resistivity. We also found that the effective resistivity, albeit for unrealistic mass ratio, still is as large as $10^4$ times the classical Spitzer value." }, "0808/0808.1722_arXiv.txt": { "abstract": "A class of very energetic supernovae (hypernovae) is associated with long gamma-ray bursts, in particular with a less energetic but more frequent population of gamma-ray bursts. Hypernovae also appear to be associated with mildly relativistic jets or outflows, even in the absence of gamma-ray bursts. Here we consider radiation from charged particles accelerated in such mildly relativistic outflows with kinetic energies of $\\sim$10$^{50}$ erg. The radiation processes of the primarily accelerated electrons considered are synchrotron radiation and inverse-Compton scattering of synchrotron photons (synchrotron self-Compton; SSC) and of supernova photons (external inverse-Compton; EIC). In the soft X-ray regime, both the SSC and EIC flux can be the dominant component, but due to their very different spectral shapes it should be easy to distinguish between them. When the fraction of the kinetic energy going into the electrons ($\\epsilon_e$) is large, the SSC is expected to be important; otherwise the EIC will dominate. The EIC flux is quite high, almost independently of $\\epsilon_e$, providing a good target for X-ray telescopes such as {\\it XMM-Newton} and {\\it Chandra}. In the GeV gamma-ray regime, the EIC would be the dominant radiation process and the {\\it Gamma-ray Large Area Space Telescope (GLAST)} should be able to probe the value of $\\epsilon_e$, the spectrum of the electrons, and their maximum acceleration energy. Accelerated protons also lead to photon radiation through the secondary electrons produced by the photopion and photopair processes. We find that over a significant range of parameters the proton component is generally less prominent than the primary electron component. We discuss the prospects for the detection of the X-ray and GeV signatures of the mildly relativistic outflow of hypernovae. ", "introduction": "\\label{sec:Introduction} While it is recognized that long gamma-ray bursts (GRBs) are associated with very energetic supernovae, sometime referred to as hypernovae, our understanding of the physics that drives and connects these events is far from being established. Observationally, long GRBs are more complicated than were thought before. For example, they appear to have subgroup that includes GRB 980425 \\citep{Galama1998}, GRB 031203 \\citep{Malesani2004,Soderberg2004} and GRB 060218 \\citep{Campana2006,Cobb2006,Pian2006}. These GRBs occurred relatively nearby and their energies radiated by prompt gamma rays were significantly smaller than the other long GRBs, and they were associated with well-studied hypernovae (SN 1998bw, SN 2003lw and SN 2006aj, respectively). Radio observations of these events \\citep{Kulkarni1998,Soderberg2006} suggest the presence of mildly relativistic ejecta, which is a different component from the usual nonrelativistic component of the supernova explosion. Their rate of occurrence may be an order of magnitude higher than that of the more energetic GRBs \\citep{Liang2007,Soderberg2006}. See also, e.g., \\citet*{Liang2006,Murase2006,Toma2007,Waxman2007,Gupta2007} for other followup studies. Very recently, a mildly relativistic outflow component has also been inferred in a supernova of type Ibc unassociated with a GRB, SN 2008D \\citep{Soderberg2008}. In this paper, we investigate the broadband radiation from the mildly relativistic ejecta associated with hypernovae, focusing especially on the high-energy photon emission in the X-ray and gamma-ray ranges. We mainly study the radiation from relativistic electrons which are primarily accelerated in shocks, but we also consider the radiation from secondary electrons which are generated by interactions involving accelerated protons. The latter were extensively studied by \\citet{Asano2008} in connection with the origin of Galactic cosmic rays \\citep{Wang2007,Budnik2008}. Our treatment of the proton component includes also the previously neglected effect of photopair production. We show that the primary electron component photon signature generally dominates over the proton component, for a wide range of energies. In particular the inverse-Compton scattering of hypernova photons due to the primary electrons appears the most promising channel for detection in the X-ray and gamma-ray ranges. The paper is organized as follows. In \\S~\\ref{sec:Spectrum of Primary Electrons}, we introduce the relevant supernova parameters and the spectrum of the primary accelerated electrons. We present our main results on the radiation from the primary electrons in \\S~\\ref{sec:Radiation from Primary Electrons}, and discuss its dependence on the parameters. Section~\\ref{sec:Radiation from secondary electrons and proton acceleration} is devoted to a discussion of the photon radiation of a proton origin and its comparison to the electron component. We summarize our conclusions in \\S~\\ref{sec:Conclusions}. ", "conclusions": "\\label{sec:Conclusions} We have studied the radiation from the mildly relativistic ejecta associated with hypernovae. The energy associated with this portion of the ejecta, about $10^{50}$ erg, starts to dissipate at radii of about 10$^{16}$ cm, where electrons and protons are shock accelerated. The electrons radiate photons via synchrotron, synchrotron-self Compton (SSC) of synchrotron photons, and external inverse-Compton (EIC) scattering of hypernova photons. The radiation is most prominent in the X-ray and GeV ranges, detectable for sources at $d < 100$ Mpc. We also studied the radiation from second and third generations of electrons and positrons that are associated with accelerated protons. The interactions that produce these leptons are the photopion and Bethe-Heitler (BH) processes, and both of these would give fluxes above the sensitivity limit of modern X-ray telescopes. However, we find that both of these components are hidden by a large flux due to the primary electron population mentioned above, for most realistic combinations of relevant parameters. Values of $\\epsilon_e \\ll 10^{-3}$ or very long lasting observations may improve the detection prospects of this component. The most promising energy range for detecting the electron SSC and EIC emission components is in the soft X-ray range. If both $\\epsilon_e$ and $\\epsilon_B$ are reasonably large, the SSC component dominates, while otherwise the EIC does (Fig.~\\ref{fig:F1keV}). The regenerated emission component from electron-positron pairs produced by $\\gamma\\gamma$ absorption of $\\gtrsim 10^2$ GeV photons is found to be always negligible. A robust feature of the EIC emission is that it provides fairly large flux in the X-ray regime, for most combinations of the parameters $\\epsilon_e$, $\\epsilon_B$ and $p$, if the supernova occurs at distances $d \\leq 100$ Mpc. When the two components are comparable, it may be easy to distinguish them through their very different spectral shapes. The GeV flux is dominated by the EIC component. Observations with {\\it GLAST} would make it possible to measure $p$ independently, if $\\epsilon_e$ is large enough, and wold also be able to probe for the spectral cutoff corresponding to the maximum acceleration energy of the radiating electrons. Finally, we briefly comment on the dependence of the results on the progenitor wind mass-loss rate. This is interesting in particular because \\citet{Campana2006} estimated a substantial mass-loss rate $\\dot M \\sim 10^{-4} M_{\\sun}$ for the progenitor of GRB 060218/SN 2006aj, which is an order of magnitude larger than the nominal value adopted here. Such a higher mass loss rate decreases the dissipation radius $R$ and the dynamical time scale $t_{\\rm dyn}$ by an order of magnitude. Even if we decrease the supernova luminosity to $L_{\\rm SN} = 10^{42}$ erg s$^{-1}$, an order of magnitude below the nominal value adopted above, while keeping the other parameters the same, we find that both the SSC and EIC photon fluxes increase by about 2--3 orders of magnitude. This more than offsets the decrease by a factor 3 of the sensitivities of X-ray telescopes caused by the smaller $t_{\\rm dyn}$, implying a very significant further improvement in the prospects for detection." }, "0808/0808.0104_arXiv.txt": { "abstract": "We examine the implications of recent measurements of the Milky Way (MW) rotation for the trajectory of the Large Magellanic Cloud (LMC). The $\\sim 14\\pm 6\\%$ increase in the MW circular velocity relative to the IAU standard of $220~\\kms$ changes the qualitative nature of the inferred LMC orbit. Instead of the LMC being gravitationally unbound, as has been suggested based on a recent measurement of its proper motion, we find that the past orbit of the LMC is naturally confined within the virial boundary of the MW. The orbit is not as tightly bound as in models derived before the LMC proper motion was measured. ", "introduction": "\\label{sec:intro} Recently, Kallivayalil et al. (2006; hereafter K06) measured the proper motion of the the Large Magellanic Cloud (LMC), and pioneered a first assessment of its 3D velocity vector. Based on this measurement, Besla et al. (2007; hereafter B07) concluded that the LMC was unlikely to have passed near the Milky Way (MW) before and was most likely formed outside of the boundaries of the Galaxy, in contrast to traditional scenarios (see references in van der Marel et al. 2002; hereafter vdM02). This new conclusion is intriguing given that there are no other examples of massive gas-rich galaxies like the LMC within the much bigger volume that separates the MW from its neighboring galaxy, Andromeda (M31). Other recent studies have also found the B07 results difficult to accept and propose alternate strategies of binding the LMC and MW either through MOND gravity (Wu et al. 2008) or by giving the LMC and its smaller counterpart, the Small Magellanic Cloud, a common halo (Bekki 2008). A couple of years after K06 published their findings, Piatek et al. (2008; hereafter P08) confirmed independently their results to within one standard deviation, although at the lower end of the inferred range of values. Also, within the past five years, the circular velocity of the MW and the distance between the Sun and the galactic center have been updated by Reid \\& Brunthaler (2004; hereafter RB04) and Gillessen et al. (2008; hereafter GGE08), respectively. These increased the likely circular velocity of the MW from the IAU standard of $V_{circ}=220$ \\kms~to 251$\\plm15$ \\kms. A value of 220 \\kms~now corresponds to a reduction of the best-fit value by two standard deviations (and equivalent to moving the Sun a total of 1 kpc closer to the center of the Galaxy). Uemura et al. (2000) used parallax measurements from Hipparcos and SKYMAP to obtain a similar value of $V_{circ}=255\\plm8$ \\kms. The rotation speed of the MW affects the analysis of the past LMC orbit in two ways. First, because the proper motion of the LMC is measured relative to the solar system which orbits the Galaxy, it is necessary to know the rotational velocity of the Sun in order to transform to the Galactocentric frame (see, e.g. vdM02). Second, the depth of the gravitational potential well of the MW (involving its estimated mass and scale radius) depends on the normalization of its rotation curve. B07 adopted the IAU standard in their analysis instead of the modified values for the Milky Way's rotation. In this {\\it Letter}, we examine the implications of the change in the MW parameters for the LMC orbit (also with the updated P08 value). In the particular geometry of the LMC orbit, both of the above-mentioned effects make the LMC more gravitationally bound to the MW owing to an increase in $V_{circ}$. Despite the small fractional magnitude of the correction in circular velocity ($\\sim 14\\pm 6\\%$), we find that the qualitative nature of the LMC orbit changes. Instead of the LMC being possibly unbound as suggested by B07, the $\\sim 10\\%$ decrease in the LMC velocity and the $\\sim 50\\%$ increase in the MW mass lead us to conclude that the LMC's past trajectory was probably confined within the virial radius of the MW. The apogalacticon distance of the orbit is comparable to the MW virial radius, as expected for a satellite that had formed at the outer edge of the Galactic halo. Traditional studies of the LMC's orbit around the MW (e.g., Murai \\& Fujimoto 1980; Lin \\& Lynden-Bell 1982; Gardiner et al. 1994; Lin et al. 1995; Gardiner \\& Noguchi 1996; vdM02; Bekki \\& Chiba 2005; Mastropietro et al. 2005; Connors et al. 2006) considered the MW as an isolated galaxy. While this might have been acceptable for studies where the LMC's orbit was confined well within the MW halo, the high LMC velocity measured by K06 and P08 implies (B07) that the apogalacticon could extend beyond the edge of the MW's halo -- where the gravitational influence of M31 is non-negligible (see Table 1 in B07). Thus, we also include the tidal effect of M31 in our calculations. In \\S \\ref{sec:method} we describe our adopted model for the mass distribution of the MW halo. We then calculate the past LMC orbit (\\S \\ref{sec:const}) and the effect of M31 on it (\\S \\ref{sec:M31}). Finally, we discuss the implications of our results in \\S \\ref{sec:disc}. ", "conclusions": "\\label{sec:disc} We have found that the $\\sim 14\\pm 6\\%$ increase in the MW circular velocity, relative to the IAU standard of $220~\\kms$, allows the LMC to be gravitationally bound to the MW. Despite its relatively high proper motion (K06; P08), the orbit of the LMC remains confined within the virial radius of the MW. Since the MW and M31 galaxies account for most of the mass in the LG, the LG itself is not much more spatially extended than the two are individually, having an estimated zero energy surface at a radius of $\\sim0.9~\\Mpc$~(Karachentsev et al. 2008). If the LMC had followed the path dictated by the P08(220~\\kms) parameters in Table 2, it would have originated from a distance of $\\ga 1.5$ \\Mpc~ away from the MW center (see Figure \\ref{fig:2dpathlg}). In comparison, if we admit the P08(251~\\kms) parameters, then the LMC originated roughly at the virial radius of the MW halo and well within the boundaries of the LG. The path suggested by P08(236~\\kms) puts the LMC origin outside the LG but closer than the P08(220~\\kms) case. If the P08(265~\\kms) model is to be believed, the LMC is on its third pass by the MW center and also originated on the edge of the MW halo. The Galactocentric distance on which this result is based ($\\rsun=9.0$~kpc) is 1--$\\sigma$ above the most recent estimate (8.4\\plm0.5 \\kpc; see GGE08), so it is not excluded. We note that during the long time-scale between perigee passes ($\\sim6$ Gyr, see Fig. \\ref{fig:distance}), the MW halo mass could evolve by tens of percent (Diemand, Kuhlen \\& Madau 2007). Future studies might use cosmological simulations to incorporate the evolution of both the host (MW-like) galaxy and its most massive (LMC-like) satellite to get a statistical understanding of their likely past interaction. Our most likely model has the LMC forming at roughly the MW virial radius, as expected from a hierarchical formation of the MW halo (Springel et al. 2008, Figs. 11 \\& 12). Also, all our quoted uncertainty for the LMC orbit stem from the uncertainty in \\rsun (and therefore \\vcirc). Errors in the proper motion were not taken into account. The K06 measurements, for example, are slightly over 1--$\\sigma$ off from the P08 measurements of the LMC proper motion. The difference between these two cases (as seen in Figure \\ref{fig:distance2}) is drastic -- in one case the LMC is clearly bound to the MW (P08), and in the other it is not (K06). A 1--$\\sigma$ shift in the other direction, however, would alter the path of the LMC into an even tighter orbit, similar to the P08(265 \\kms) model. With the value of the distance between the Sun and the Galactic center at its IAU value (GGE08, KLB86), $\\sim$8.5$\\plm.5~\\kpc$, the total mass of the MW Galaxy increases by a factor of 1.23--1.75 relative to the values inferred for the lower distance of 8\\plm.5 \\kpc ~(Reid 1993). B07 correctly rules out a larger mass for the MW but only considers the low mass ($1\\times10^{12}\\msun$) and the high mass ($2\\times10^{12}\\msun$) models of Klypin et al. (2002) which bracket our preferred range. Xue et al. (2008) describe the uncertainties of the prior assumption that the galactic \\vcirc~ is 220 \\kms. Table \\ref{tab:MW} provides the corresponding range of masses for the MW halo, all calculated from observational data, fit either to a flat rotation curve or an NFW profile, such as the one we have adopted in this work. \\begin{table} \\caption{Milky Way Halo Mass in Recent Studies} \\label{tab:MW} \\begin{tabular}{@{}lcc} \\hline Author & Mass & Model \\\\ ~ & ($10^{12}\\msun$) & ~ \\\\ \\hline \\hline Wilkinson \\& Evans (1999)& $1.9^{+3.6}_{-1.7}$ & FRC$^{1}$ \\\\ Sakamoto, Chiba \\& Beers (2003) & $2.5^{+0.5}_{-1.0}$ & FRC \\\\ Sakamoto, Chiba \\& Beers ({\\it w/o} Leo I) & $1.8^{+.04}_{-0.7}$ & FRC \\\\ Smith et al. (2007) & $1.42^{+1.14}_{-0.54}$ & NFW \\\\ Xue et al. (2008) & $0.79\\plm0.15$ - & NFW\\\\ ~ & ~~~$1.18\\pm0.28$ & \\\\ Li \\& White (2008)& $2.43$ & N-Body$^{2}$\\\\ \\hline \\end{tabular} $^{1}$FRC denotes a Flat Rotation Curve, as in an isothermal sphere. These mass calculations have cutoffs of $\\sim 50~\\kpc$. $^{2}$e.g. the Millennium simulation. Here the limiting radius is the virial radius, similar to the NFW calculation. \\end{table} The $\\sim 50 \\%$ increase in mass (from $1\\times10^{12}~\\msun$ to $1.485\\times10^{12}~\\msun$) and the $\\sim 6\\%$ decrease in the velocity of the LMC (P08(220~\\kms) to P08(251~\\kms)) have a comparable influence on making LMC orbit bound to the MW. The combined effect of these changes is not equivalent to a change in the MW mass, as considered by B07. The K06(251~\\kms) and P08(220~\\kms) lines in Figures \\ref{fig:distance2} and \\ref{fig:distance} correspond to roughly the same LMC velocity (as derived from the IAU standard with the P08 proper motion) but a different MW mass. The higher mass of the MW contributes about half of the overall difference between these cases. This parallels the comparison made in B07 between the K06(220~\\kms) and the K06(220~\\kms)$+4\\sigma$ numbers in their {\\it Fiducial} and {\\it High Mass} Models, except that our findings do not require a $4-\\sigma$ deviation from the measured LMC velocity or a $2-\\sigma$ deviation from the mass of the MW in order to make the LMC orbit bound to the MW. The previous suggestion (B07) of possibly disqualifying the LMC and SMC as MW satellites has undesirable implications for the satellites in the MW halo. M31 has 18 satellites, 5 of which are gas rich, whereas the MW, if the LMC and SMC are no longer bound to it, has only 12 bound satellites, 2 of which are gas rich (Karachentsev 2005, vdMG08). M31's satellites range in mass from 0.58 to 500 $\\times10^{8}\\msun$, whereas the remaining MW satellites are in the range of 0.1--1$\\times10^{8}\\msun$ (Mateo 1998). Excluding the LMC and SMC as satellites of the MW appears unnatural, as there is no reason for the two comparably sized galaxies to have a large disparity in the abundance of massive satellites. Moreover, the chance of finding massive galaxies like the LMC and the SMC so close to the MW center requires a special coincidence if they are unbound to the MW, since they would have spent most of their orbital time far away from the MW in that case. Yet, no similar galaxies are known to exist in the much larger volume between M31 and the MW. Although these statistical arguments are not definitive, they support indirectly the higher updated values of \\vcirc (RB04) and \\rsun (GGE08) for the MW. No work on the Magellanic Clouds would be complete without a mention of the Magellanic Stream (MS). There is $<2\\%$ change in the trajectory over the length of the MS (100 degrees) between the P08(220\\kms) and the P08(251\\kms) models, both of which are well within the error margins in Figure 8 of B07. We reiterate the concerns presented in B07 that neither tidal stripping nor ram pressure can fully explain the orientation of the Stream. \\noindent {\\bf Acknowledgments.} We thank Gurtina Besla and Mark Reid for useful discussions. This work is supported in part by NASA grant NNX08AL43G, by FQXi, and by Harvard University and Smithsonian Astrophysical Observatory funds." }, "0808/0808.0332_arXiv.txt": { "abstract": "While the particle hypothesis for dark matter may be very soon investigated at the LHC, and as the PAMELA and GLAST satellites are currently taking new data on charged and gamma cosmic rays, the need of controlling the theoretical uncertainties affecting the possible indirect signatures of dark matter annihilation is of paramount importance. The uncertainties which originate from the dark matter distribution are difficult to estimate because current astrophysical observations provide rather weak dynamical constraints, and because, according to cosmological N-body simulations, dark matter is neither smoothly nor spherically distributed in galactic halos. Some previous studies made use of N-body simulations to compute the $\\gamma$-ray flux from dark matter annihilation, but such a work has never been performed for the antimatter (positron and antiproton) primary fluxes, for which transport processes complicate the calculations. We take advantage of the galaxy-like 3D dark matter map extracted from the HORIZON Project results to calculate the positron and antiproton fluxes from dark matter annihilation, in a model-independent approach as well as for dark matter particle benchmarks relevant at the LHC scale (from supersymmetric and extra-dimensional theories). We find that the flux uncertainties arise mainly from fluctuations of the local dark matter density, and are of $\\sim$ 1 order of magnitude. We compare our results to analytic descriptions of the dark matter halo, showing how the latter can well reproduce the former. The overal antimatter predictions associated with our benchmark models are shown to lie far below the existing measurements, and in particular that of the positron fraction recently reported by PAMELA, and far below the background predictions as well. Finally, we stress the limits of the use of an N-body framework in this context. ", "introduction": "\\label{sec:intro} The idea that dark matter is made of exotic weakly interacting massive particles (WIMP) is very appealing in the sense that the related existing frameworks in particle physics beyond the standard model (BSM) could solve (i) problems inherent to elementary particle theory, e.g. the force unification scheme and/or the stabilization of the theory against quantum corrections, which would thus confer a more fundamental meaning to this theory; and in the sametime (ii) the dark matter issue as characterized in astrophysics and cosmology (see e.g.~\\cite{review_dm_murayama_07} for a recent review). Indeed, the cold dark matter (CDM) paradigm seems to feature a powerful and self-consistent theory of structure formation, which at present is able to reproduce and explain most of large and mid scale observations (see e.g.~\\cite{2007NuPhS.173....1P} and references therein). While the Large Hadron Collider at CERN (LHC) is about to start hunting BSM particle physics and may soon provide new insights about the dark matter particle hypothesis, high energy astrophysics experiments are also of strong interest to try to determine the actual nature of dark matter. Indeed, if dark matter is made of self-annihilating particles, a property which is found in many BSM models and which provides a natural mechanism to explain the dark matter cosmological abundance as observed today, there should be traces of annihilations imprinting the cosmic ray spectra (indirect detection of dark matter, see e.g.~\\cite{susy_dm_jungman_etal_96,review_dm_bergstrom_00, review_dm_bertone_etal_05,review_dm_carr_etal_06}). Therefore, satellite experiments such as PAMELA and GLAST, which are dedicated to large field of view observations of charged cosmic rays and $\\gamma$-rays, respectively, in the MeV-TeV energy range, should be able to yield additional constraints or smoking guns very soon in this research field~\\cite{2006JCAP...12..003M}. The most promising astrophysical messengers that could trace the annihilation processes are indeed gamma-rays and antimatter cosmic rays, which have been first investigated in~\\cite{1978ApJ...223.1015G} and in~\\cite{1984PhRvL..53..624S} respectively. Confinement and diffusion on magnetic turbulences limit the origin of the latter to our Galaxy, whereas the former, which only experiences the common $r^{-2}$ flux dilution, may be observed even when emitted from extra-galactic regions. Nevertheless, predictions of these exotic signals are affected by many uncertainties coming from (i) the underlying WIMP model (ii) the distribution of dark matter in the relevant sources (iii) the propagation of charged cosmic rays in the Galaxy in the case of searches in the antimatter cosmic ray spectra. The first point is encoded in the WIMP mass, the annihilation cross-section and the annihilation final states which fully define the injected cosmic ray spectra: this has been widely investigated for years. The second point is currently surveyed by the state-of-the-art N-body experiments, but the use of N-body dark matter maps in the context of indirect detection has only been performed for $\\gamma$-ray predictions~\\cite{2003MNRAS.345.1313S,2007ApJ...657..262D,2008arXiv0801.4673A, 2008arXiv0805.4416K}. Besides, there has been lots of efforts to estimate the effect of varying the dark matter distribution by means of analytical calculations, allowing for instance to study the effects of sub-halos on the $\\gamma$-ray~\\cite{1999PhRvD..59d3506B, 2002PhRvD..66l3502U,berezinsky_etal_03,berezinsky_etal_06, 2008MNRAS.384.1627P} as well as on the antimatter primary fluxes~\\cite{2007A&A...462..827L,2008A&A...479..427L}. Such analytical studies mostly rely on the statistical information supplied by N-body simulations. However, though they provide very nice theoretical bases to unveil the salient effects of any change in the relevant parameter space, and permit to go beyond the current numerical resolution limits, they are usually bound to simplifying hypotheses, such as the sphericity of sources. Especially, no study of the antimatter signatures has been directly performed in the frame of N-body environments up to now. Finally, point (iii), referring to cosmic ray propagation, has also widely been investigated for antiprotons~\\cite{2004PhRvD..69f3501D,2005JCAP...09..010L} and positrons~\\cite{2008A&A...479..427L,2008PhRvD..77f3527D}, and while the transport processes depend on sets of still degenerate parameters within different propagation models, the origins of uncertainties are now rather well understood. In this paper, we make use of a realistic (in the sense of \\lcdm\\ cosmology) 3D dark matter distribution of a galactic-sized halo extracted from the HORIZON Project simulation results to study the antimatter signals. Our main purposes are to characterize and quantify the uncertainties coming from dark matter inhomogeneities (not necessarily clumps) and departure from spherical symmetry in a virialized Milky-Way-like object; and finally make predictions for different dark matter particle candidates. Despite the rather low resolution of our numerical data compared with the highest-level artillery on the market, as portrayed, e.g., by the Via Lactea simulations~\\cite{2007ApJ...657..262D,2008arXiv0805.1244D}, we study for the first time the production of antimatter cosmic rays from dark matter annihilation in an N-body framework. The resolution issue is however less important for charged cosmic rays than for gamma-rays, because the signal is strongly diluted by diffusion effects, which therefore tends to smooth the yields from high density fluctuations, provided the latter are not too close to the observer. Nevertheless, as very local density fluctuations are expected to affect the antimatter signals, we will quantify and correct the consequence of loss of resolution by analytically extrapolating our results. For the WIMP candidates, we will use a model-independent approach as well as typical models of BSM particle theories, some of them being observable at the LHC. This article is parted as follows: we recall the salient features of the HORIZON N-body experiment in Sect.~\\ref{sec:horizon}; then, in Sect.~\\ref{sec:crs}, we briefly review the positron and antiproton propagation before sketching the method we adopt to connect the propagation to the N-body source terms; we describe the WIMP models in Sect.~\\ref{sec:wimp}; we finally present and discuss our results and predictions of the primary positron and antiproton fluxes in Sect.~\\ref{sec:res} (where the expert reader is invited to go directly) before concluding. ", "conclusions": "\\label{sec:concl} In this paper, we have studied the theoretical uncertainties associated with dark matter density fluctuations affecting the primary antimatter cosmic ray fluxes (positrons and antiprotons), possibly produced by dark matter annihilation in the Galaxy, by using an N-body simulation together with analytical descriptions of a galaxy-like dark matter halo. This is the first attempt in this research field to directly connect the source term of the cosmic ray propagation equation to a 3D density map coming from a cosmological N-body framework, while this has already often been used for predictions of gamma-ray fluxes. Indeed, the latter situation does not involve cosmic ray propagation, which seriously complicates the calculations, and makes them much more time consuming. The main purposes of this work was to quantify those uncertainties, and in some cases to try to understand them by analytical means. While this was achieved in a WIMP model-independent setting, we also aimed at reviewing the status of predictions for many different well motivated dark matter candidates in particle physics, supersymmetric or not. The N-body framework led us to study several effects, such as non-spherical dark matter distributions (Sect.~\\ref{subsubsec:ellipt}) and density fluctuations (Sect.~\\ref{subsubsec:inhom}), which are not accounted for when using the traditional spherical smooth halo models. We have shown that, as expected, fluxes are mainly affected at cosmic ray energies corresponding to short propagation characteristic lengths (low/high for antiprotons/positrons) and to local contributions, while when looking at energies of large propagation lengths, the smoothing due to diffusion erases the peculiarities in the spectra, and is well reproduced by a spherical halo model. The main uncertainties on the predicted fluxes come therefore from those on the local dark matter environment, which we have shown to fluctuate a lot in the N-body framework. This can lead to $\\pm$ 1 order of magnitude in terms of flux, at high energy for positrons, and low energy for antiprotons. This could be more constrained with much better limits on the local dark matter density. Moreover, though our simulation has a too poor spatial resolution to resolve substructures inside the galactic halo, we have extrapolated our results in order to include the potential effects of the presence of sub-halos by using the analytical method detailed in~\\cite{2008A&A...479..427L}. We have considered two situations: one \\emph{maximal}, analytically extending the clump mass spectrum down to $10^{-6}\\Msun$, led to a flux enhancement by an energy-dependent factor reaching $\\sim 4$ at maximum; a second, \\emph{Via-Lactea-like}, involved clumps down to $10^6\\Msun$ only and did not result in any flux enhancement. Consequently to the previous points, the smooth and spherical modeling of the dark matter halo leads to a very good approximation of a more complete and complex situation, as derived from an N-body simulation, at large characteristic propagation lengths for antimatter cosmic rays (low/high energies for positrons/antiprotons). Nevertheless, though predictions in the spectral regions associated with those large characteristic scales are also the less affected by the uncertainties coming from dark matter density fluctuations, we remind that they are still very sensitive to the used propagation model. Beside this trial for directly \\emph{measuring} the theoretical uncertainties with an N-body experiment, we reviewed the predictions of the positron and antiproton primary contribution associated with some specific particle physics dark matter candidates (Sect.~\\ref{subsec:dm_models}). We have not only included some popular supersymmetric and extra-dimensional models, but also some more specific ones based on minimality arguments, such as the little Higgs model, or the inert doublet model. We have shown that predictions are well lower than the existing data, except perhaps for the antiproton flux associated with the lightest neutralino model, just because of the more favorable $1/\\mchi^2$ factor appearing in the flux expression. Nevertheless, even if the latter appears in tension with the current data, its contribution to the antiproton flux occurs at energies where solar modulation effects are expected to be important, and it would be hard to find a clear signature in this spectral region. The other models are well below the experimental measures, but higher energy measurements could perhaps unveil some spectral transitions followed by an energy cut-off which could possibly be attributed to dark matter. In this case, uncertainties associated with the dark matter density fluctuations would be much less stringent for the antiproton signal than for the positron because of propagation scale arguments. On the contrary, a nearby clump would favor a detection with positrons at high energy, much more than with antiprotons. Anyway, higher energy data are necessary in this field, for these antimatter species as well as for standard nuclei species which will yield much better constraints to the propagation modeling. Obviously, indications of a WIMP candidate mass, as expected from the LHC, would provide the relevant spectral region where to concentrate the astrophysical searches, and would allow to concentrate on much more subtle effects. Finally, we made a self-criticism exercise -- Sect.~\\ref{subsec:dm_limits} -- by comparing our different halo models to the star kinematic data available for our Milky-Way Galaxy, after subtraction of the baryonic contribution as modeled by~\\cite{2006CeMDA..94..369E}. We show that the dark matter contribution to the velocity field as inferred from our N-body simulation, as well as more general ones, are far to reproduce the kinematic data, systematically leading to a mass excess in the central regions ($\\lesssim 4$ kpc) of the Galaxy. This clearly calls for restraint in our potential claims, and clearly points towards some large amount of ignorance in the intimate (and gravitational) relations between dark matter and baryons." }, "0808/0808.2047_arXiv.txt": { "abstract": "We provide a new derivation of the anisotropies of the cosmic microwave background (CMB), and find an exact expression that can be readily expanded perturbatively. Close attention is paid to gauge issues, with the motivation to examine the effect of super-Hubble modes on the CMB. We calculate a transfer function that encodes the behaviour of the dipole, and examine its long-wavelength behaviour. We show that contributions to the dipole from adiabatic super-Hubble modes are strongly suppressed, even in the presence of a cosmological constant, contrary to claims in the literature. We also introduce a naturally defined CMB monopole, which exhibits closely analogous long-wavelength behaviour. We discuss the geometrical origin of this super-Hubble suppression, pointing out that it is a simple reflection of adiabaticity, and hence argue that it will occur regardless of the matter content. ", "introduction": "The anisotropies in the cosmic microwave background (CMB) reveal a great deal about our Universe, since they persist essentially unscathed from the epoch when fluctuations were well described by simple linear theory. The comparison of CMB observations with theory has become a mature subject, and has played an important role in forming our current understanding of the Universe (see, \\eg, \\cite{wmap5}). Critical to that comparison is the accurate theoretical calculation of the anisotropies. Since the pioneering work of Sachs and Wolfe \\cite{sw67}, the theoretical anisotropies have been refined and recalculated using different formalisms many times (see, \\eg, \\cite{panek86,magueijo92,rsxd93,wh97,dunsby97,cl98,hn99}). Accurate calculations are now readily available via public code packages such as \\textsc{camb} \\cite{lcl00,notecamb}. In the present work, we revisit the calculation of anisotropies. While our results may not lead to more accurate or efficient calculations, we hope that they will help to clarify some of the conceptual issues surrounding the calculations. In particular, our approach makes explicit the physical meaning of the various contributions to the anisotropies. Crucial to this is our use of the covariant approach to cosmology (see \\cite{ee98,tcm08} for reviews), which is ideal for writing exact solutions and for physical clarity. We present a remarkably simple but exact expression for the anisotropy, which applies to arbitrary spacetimes and includes the effects of tensor as well as scalar \\perts\\ and any line-of-sight integrated Sachs-Wolfe (ISW) effect. This general result can be readily expanded perturbatively, and here we turn to the metric formalism for computational efficiency and show that we recover previous results in the literature. A main motivation for our work is in examining the behaviour of the anisotropies due to super-Hubble fluctuations (we use the terms ``super-Hubble'', ``long-wavelength'', and ``large-scale'' interchangeably, to mean scales larger than the current Hubble or last scattering radius), where gauge issues are paramount. This question has been examined before in the context of the Grishchuck-Zel'dovich effect \\cite{gz78}, which describes the large-angular-scale anisotropies that result from super-Hubble modes. In the context of a matter-dominated universe with adiabatic fluctuations, it was shown that the CMB dipole receives strongly suppressed contributions from long-wavelength modes. A claim was made in Ref.~\\cite{turner91} that this suppression would not occur in models with cosmological constant, so that we could ``see'' very long-wavelength structure in the dipole. It was also found that in the presence of {\\em isocurvature} \\perts, the suppression may not occur (see \\cite{langlois96} and references therein). To study this issue, we construct a transfer function that describes the scale dependence of contributions to the dipole. Working by analogy, we carefully define a CMB {\\em monopole} \\pert, and find its transfer function. As is well known, such a monopole cannot be observable, but we show that its {\\em variance} is well defined theoretically. Our definitions have simple interpretations: the dipole measures the departure of radiation and matter comoving worldlines, while the monopole measures how well radiation and matter constant-density hypersurfaces coincide. The usefulness of the monopole will be in examining its long-wavelength behaviour, where it will help to clarify the dipole case. We show that the contributions to both dipole and monopole vanish for large scale sources, even in the presence of a cosmological constant. We close by pointing out that this is a direct consequence of adiabaticity, and hence that this result is expected to hold regardless of the matter content of the Universe, unless isocurvature modes are present. Another potential reason that a careful treatment of the monopole may be of interest involves the measurement of the mean CMB temperature and its relation to constraints on other cosmological parameters. The mean temperature is currently measured to a precision of a few parts in $10^4$ \\cite{mfsmw99}. It has been pointed out that the precision of this measurement could be improved by nearly two orders of magnitude with currently available technology \\cite{fm02}. Such a measurement would reach the naive cosmic variance limit of a part in $10^5$ as suggested by the observed amplitude of fluctuations (see \\cite{wz08} for a related discussion). It would then become necessary to be very careful about exactly what information the mean temperature measurement is giving us, and about the nature of monopole fluctuations. Relevantly, recent studies have examined the importance of the mean temperature measurement to our ability to constrain the cosmological parameters \\cite{cs08,hw08}. Since the covariant approach to cosmology is essential to this work, we begin in Sec.~\\ref{covarsec} with a summary of the required formalism. Next, in Sec.~\\ref{swsec} we present the derivation of the Sachs-Wolfe effect, beginning with an exact result before specializing to first order and recovering previous results. In Sec.~\\ref{secmonodi} we present calculations of the dipole and monopole transfer functions, and we examine their long-wavelength behaviour in Sec.~\\ref{seclongwl}. Finally we discuss our results in Sec.~\\ref{discusssec}. The Appendices summarize relevant material in the metric formalism, and demonstrate both the gauge invariance and the gauge dependence of our results. We use signature $(-,+,+,+)$, and greek indices indicate four-tensors, while latin indices indicate spatial three-tensors. ", "conclusions": "\\label{discusssec} Our result in Sec.~\\ref{seclongwl} that the dipole and monopole receive suppressed contributions from large scales, even in the presence of a dominant cosmological constant, strongly suggests that the cancellations involved are not accidental. To understand the origin of this behaviour, consider first the monopole case. Physically, the suppression as $k \\ra 0$ in the monopole transfer function shown in Figs.~\\ref{transfnc} and \\ref{monotrans} means that surfaces of uniform total and radiation energy density coincide on the largest scales, according to our definition of the monopole in Sec.~\\ref{secmono}. But this is just the statement of adiabaticity: an adiabatic matter-radiation fluid is characterized by the condition \\beq \\fr{\\del\\rho}{\\dot\\rho} = \\fr{\\del\\rho_{(\\gam)}}{\\dot\\rho_{(\\gam)}}, \\label{adcond} \\eeq on the largest scales. Therefore, \\eq(\\ref{uedgtrns}) shows that the same gauge transformation takes us to both constant total matter and constant radiation \\hsfs; in other words, those surfaces must coincide. It is important to point out that adiabatic long-wavelength modes remain adiabatic under evolution \\cite{mfb92}, and indeed the comoving curvature \\pert\\ $\\R$ remains constant in time (see, \\eg, \\cite{ll00}). This trivial evolution on super-Hubble scales means that when constant matter and radiation density surfaces coincide on large scales at last scattering, due to adiabatic initial conditions, they must also coincide today. An analogous situation holds for the dipole. In this case, the adiabaticity condition implies that, on large scales, the radiation and total matter comoving worldlines coincide (\\ie, there is no ``peculiar velocity'' isocurvature mode between the two components). According to our definition of the dipole in Sec.~\\ref{dipsubsec}, this is simply the statement that the dipole is suppressed on large scales. This leads us to the important conclusion that this insensitivity to long-wavelength sources must apply regardless of the matter content of the Universe, as long as adiabaticity holds. Therefore the suppression we found for the specific case of cosmological constant (or Einstein-de~Sitter) universes must in fact occur in general. Note that this may have relevance to very recent discussions regarding a potential power asymmetry in the CMB \\cite{ekc08}. The exquisite cancellations visible at large scales in Figs.~\\ref{ditrans} and \\ref{monotrans} between the SW and ISW components illustrate a previously unrecognized relation between the two, which is enforced by the condition of adiabaticity. In brief, the dipole and monopole as defined here are just measures of departures from the adiabaticity condition, \\eq(\\ref{adcond}), which generally occur on small scales. Are these results sensitive to the definitions of dipole and monopole used? As long as the dipole and monopole are defined {\\em physically}, \\ie\\ in relation to locally measureable quantities, then the results must still hold. An important example which does {\\em not} satisfy this criterion is the zero-shear, or longitudinal gauge, since it is not defined in terms of local, observable matter quantities. If the dipole (or monopole) is defined with respect to a zero-shear frame, then it may appear from theoretical calculations that the dipole {\\em is} sensitive to long-wavelength modes. However, this sensitivity cannot be observable, since zero-shear frames cannot be {\\em uniquely} constructed locally. (A linear boost or ``tilt'' of a zero-shear frame is still a zero-shear frame.) Thus great care must be taken when considering behaviour on large scales with longitudinal gauge. Indeed, another way to understand this result is to realize that, in the limit $k/(aH) \\ra 0$, an adiabatic \\pert\\ mode locally becomes essentially pure gauge, and can be removed within any {\\em sub-Hubble} region by a simple boost coordinate transformation. Such a mode is indistinguishable locally from a homogeneous background, and hence cannot have any observational consequences such as a dipole anisotropy \\cite{unruh98}. Finally, we note that when the condition of adiabaticity is relaxed, then our conclusions no longer hold \\cite{langlois96}. In the presence of isocurvature \\perts, it is possible that a {\\em physically} defined dipole be sensitive to super-Hubble modes, since the extra freedom allows for a relative tilt between comoving matter and radiation comoving worldlines on large scales. While we have emphasized here the consequences of adiabaticity, it is hoped that other applications will follow from the exact formalism for CMB anisotropies that we have developed. One possibility is the evaluation of anisotropies, in particular the ISW effect, in void models of acceleration \\cite{celerier99} (see \\cite{enqvist08} for a brief review), which have not yet been confronted with observations at the perturbative level \\cite{zibin08}. Note however that the present approach is not limited to calculating CMB anisotropies, and that more generally it is applicable to calculating redshifts in arbitrary spacetimes. Potential uses include calculating the redshift-luminosity distance relation in perturbed spacetimes (see, \\eg, \\cite{hg06}). {\\em Note added:} When this work was essentially complete, a related paper appeared \\cite{eck08}, which appears to support our conclusion that long-wavelength \\perts\\ cannot effect the CMB dipole." }, "0808/0808.3088_arXiv.txt": { "abstract": "{Disk galaxies with a spheroidal component are known to host Supermassive Black Holes (SMBHs) in their center. Unequal-mass galaxy mergers have been rarely studied despite the fact that they are the large majority of merging events by number and they are associated with the typical targets of gravitational wave experiments such as LISA. We perform N-body/SPH simulations of disk galaxy mergers with mass ratios 1:4 and 1:10 at redshifts z=0 and z=3. They have the highest resolution achieved so far for merging galaxies, and include star formation and supernova feedback. Gas dissipation is found to be necessary for the pairing of SMBHs in these minor mergers. Still, 1:10 mergers with gas allow an efficient pairing only at high z when orbital times are short enough compared to the Hubble time. ", "introduction": "Observations of nearby galaxies show that Supermassive Black Holes (SMBHs) ranging between $\\sim10^6$ and $\\sim10^9 M_\\odot$ inhabit the centers of virtually all massive galactic spheroids, from massive ellipticals to pseudo-bulges of late-type galaxies. Their masses, as inferred from dynamical measurements, appear to be correlated with various properties of their hosts, e.g. bulge luminosity and mass (e.g. \\citet{kormendyrichstone95, magorrian98, marconihunt03}), velocity dispersion \\citep{tremaine02, ferrarese00}, concentration of the light profile \\citep{graham01}. In the $\\Lambda$CDM model, galaxies build up their masses hierarchically starting from initial, gravitationally amplified fluctuations (e.g. \\citet{whiterees78}); therefore, every time two galaxies merge, the remnant is expected to host two (or more) SMBHs. The formation of a binary SMBH sistem has been shown to proceed quickly once both the compact objects are embedded in a circumnuclear gaseous disk (see \\citet{lucio07}; see also Mayer, Kazantzidis \\& Escala in this volume), but whether the large scale merger can lead them to such a favorable configuration is still a matter of debate: while mergers between galaxies of equal mass have been quite extensively explored in literature and seem to lead to the formation of a SMBH pair \\citep{stelios05,springel05}, little attention has been paid to the fate of SMBHs in events which are much more frequent in the typical history of a $\\Lambda$CDM galaxy, i.e. minor mergers with mass ratios from 1:4 to 1:10 and less \\citep{stewart08}. ", "conclusions": "Understanding the formation of SMBH binaries is of fundamental importance for the search of gravitational waves as well as for all studies of black hole demography and host galaxy coevolution. SMBH pairing in unequal-mass mergers depends crucially on gasdynamical effect: satellites are disrupted too quickly, leaving wandering SMBHs in all the collisionless cases we studied. {\\it The presence of gas seems to be necessary for the pairing of SMBHs}. While it also appears sufficient for $q=0.25$, lower mass objects are more heavily affected by feedback from star formation, and their outcome could be very sensitive to the details of the physical processes involved." }, "0808/0808.0161_arXiv.txt": { "abstract": "Ultra-high energy cosmic rays (UHECRs) accelerated in the jets of active galactic nuclei can accumulate in high magnetic field, $\\sim 100$ kpc-scale regions surrounding powerful radio galaxies. Photohadronic processes { involving UHECRs and photons of the extragalactic background light} make ultra-relativistic electrons and positrons that initiate electromagnetic cascades, leading to the production of a $\\gamma$-ray synchrotron halo. We calculate the halo emission in the case of Cygnus A and show that it should be detectable with the { Fermi Gamma ray Space Telescope} and possibly detectable with ground-based $\\gamma$-ray telescopes if radio galaxies are the sources of UHECRs. ", "introduction": "Active galactic nuclei (AGN) and gamma-ray bursts are considered as two of the most plausible classes of astrophysical accelerators of extragalactic ultra-high energy cosmic rays \\citep[UHECRs; see, e.g.,][]{hh02}. The recent report of the \\citet{Auger07} about clustering of the arrival directions of UHECRs with energies $E \\gtrsim 6\\times 10^{19}$ eV within $\\approx 3^\\circ$ of the directions to AGN at distances $d\\lesssim 75\\, \\rm Mpc$ strongly suggests that effective production of cosmic rays with energies $E\\sim 10^{20}\\,\\rm eV$ takes place in at least one of these source classes. Because of photohadronic GZK interactions of protons or ions with the CMB radiation, the study of super-GZK UHECRs from sources at $d\\gtrsim 100\\,\\rm Mpc$ becomes impossible with cosmic-ray detectors like the Pierre Auger Observatory \\citep{hmr06}. Powerful AGN, as well as GRBs, are mostly located at larger distances. { Relativistic beams of energy from the central nuclei of AGN are thought to power the multi-kpc scale radio lobes of powerful galaxies and form an extended cavity \\citep{sch74}. Acceleration of UHECRs in the compact inner jets of the radio galaxy on the sub-parsec scale, followed by production of collimated beams of ultra-high energy neutrons and gamma-rays, provides a specific mechanism to transport energy to the radio lobes and cavity \\citep{ad03}.} When the neutron-decay UHECR protons interact with the extragalactic background light (EBL), which is dominated by the CMB radiation, ultra-relativistic electrons (including positrons) and $\\gamma$ rays with $E\\gtrsim 10^{18}\\,\\rm eV$ are produced. Such secondaries can initiate pair-photon cascades to form large multi-Mpc scale halos of GeV/TeV radiation due to Compton and synchrotron processes \\citep{acv94,aha02,ias05}. In this Letter, we predict that synchrotron GeV fluxes from UHECR AGN sources are detectable with the Fermi Gamma ray Space Telescope (FGST; formerly the Gamma ray Large Area Space Telescope, GLAST) if UHECRs are captured in the vicinity of radio galaxies for sufficiently long times. Magnetic fields at the $\\gtrsim \\mu$G level in the kpc -- Mpc vicinity from the AGN core are required to isotropize UHECRs accelerated by jets of radio galaxies. Indeed, for protons with energy $E\\equiv 10^{20} E_{20} \\,\\rm eV$, the mean magnetic field $B$ required to provide gyroradii smaller than size $r$ is $B\\gtrsim 10^{-4} E_{20} r_{\\,\\rm kpc}^{-1}\\,\\rm G$, where $r_{\\rm kpc} =r/1\\,\\rm kpc$ is the spatial scale where the isotropization occurs. Here we consider the specific case of the powerful radio galaxy Cygnus A, { where the mean magnetic field in the surrounding cavity could reach 10 -- 100 $\\mu$G at 100 kpc scales}. Its properties are considered in Section 2, and calculations are presented in Section 3. We summarize in Section 4. ", "conclusions": "If the radio lobes of Cyg A are powered by UHECR production from the inner pc-scale jets, then trapping of these particles in the surrounding strong magnetic-field region leads to the production of secondary $\\gamma$ rays that should be significantly detected with the FGST in one year of observation if the jet injection age is $\\gtrsim 100$ Myr. If radio galaxies are not the sources of UHECRs, then Cygnus A will not be detected by the FGST. Cyg A could also be detected with VERITAS in a 50 hour pointing, depending on the duration of activity of the central engine and the level of the EBL. Detection of GeV $\\gamma$ rays from Cyg A with the FGST might also be expected to arise from other processes. The $\\sim 10$ -- 100 GeV radio-emitting electrons from the lobes of radio galaxies will Compton-scatter CMB photons to MeV -- GeV energies \\citep[e.g.,][]{che07}. For the strong magnetic field, $\\approx 60~\\mu$G, in the lobes of Cyg A, however, the ratio of the magnetic-field to CMBR energy densities is $\\approx 400$. Thus the total energy flux of Compton-scattered CMBR from Cyg A is $\\approx 10^{45}$ erg s$^{-1}/[400(4\\pi d^2)] \\cong 4\\times 10^{-13}$ ergs cm$^{-2}$ s$^{-1}$, with the $E\\cdot F(E)$ flux a factor of $\\sim 5$ -- 10 lower. As can be seen from Fig.\\ \\ref{f2}, this process is almost two orders of magnitude below the UHECR-induced synchrotron flux, and well below the FGST sensitivity. { \\citet{ias05} considered fluxes expected from the $\\lesssim 1 \\,\\rm Mpc$ halos of clusters of galaxies with weaker magnetic fields, $B\\simeq 0.1$-$1\\,\\mu$ G in a model where acceleration of UHECRs occurs in accretion shocks in the cluster. Because of lower maximum energies of accelerated protons, $E\\lesssim 10^{19}\\,\\rm eV$, and lower magnetic fields, this model predicts hard spectral fluxes of Compton origin peaking at TeV energies}. \\citet{ga05} predicted that synchrotron radiation from $\\gtrsim 10^{18}\\,\\rm eV$ electrons is produced by secondaries of UHECRs that leave the acceleration region and travel nearly rectilinearly through weak intergalactic magnetic fields at the level $B \\sim 10^{-7}$ -- $10^{-9}\\,$G. These sources would appear as point-like quiescent GeV -- TeV sources with spectra in the GeV domain as hard as $\\alpha \\cong -1.5$, and very soft, $\\alpha \\lesssim -3$ spectra in the 100 GeV -- TeV domain. { In contrast to both these models}, we predict soft 0.1 -- 1 GeV spectra with $\\alpha \\cong -2.5$ and hard, $\\alpha \\cong -2$ spectra at TeV energies due to the much higher magnetic field in the confinement region. These models can be clearly distinguished if Cygnus A is resolved by the Fermi Gamma ray Space Telescope or the ground-based $\\gamma$-ray telescopes, as the emission region in our model subtends an angle $\\approx 10^\\prime$. Because Cygnus A lies outside the GZK horizon, only UHECRs with energy below the GZK energy could be correlated with this source. Other closer FRII radio galaxies that are correlated with the arrival directions of UHECRs are, however, candidate sources of $\\gamma$ rays made through the mechanism proposed here. IGR J21247+5058 at $z = 0.02$ or $d\\approx 80$ Mpc, recently discovered with INTEGRAL \\citep{mol07}, is 2.1 degrees away from a HiRes Stereo event with $E > 56$ EeV (C.\\ C.\\ Cheung, private communication, 2008).\\footnote{This radio galaxy was identified as such only recently and would not have appeared in the list of AGN used by the HiRes collaboration in their correlation study \\citep{abb08}.} PKS 2158-380 at $\\approx 140$ Mpc is also within 3.2$^\\circ$ degrees of an Auger UHECR with $E> 57$ EeV \\citep{mos08}. By comparison with Cyg A, these are low luminosity FRIIs, and their predicted flux level will require detailed modeling for each source, as done here for Cyg A. Variability of $\\gamma$-ray emission would rule out our model." }, "0808/0808.2813_arXiv.txt": { "abstract": "We make a case for the existence for ultra-massive black holes (UMBHs) in the Universe, but argue that there exists a likely upper limit to black hole masses of the order of $M \\sim 10^{10} \\msun$. We show that there are three strong lines of argument that predicate the existence of UMBHs: (i) expected as a natural extension of the observed black hole mass bulge luminosity relation, when extrapolated to the bulge luminosities of bright central galaxies in clusters; (ii) new predictions for the mass function of seed black holes at high redshifts predict that growth via accretion or merger-induced accretion inevitably leads to the existence of rare UMBHs at late times; (iii) the local mass function of black holes computed from the observed X-ray luminosity functions of active galactic nuclei predict the existence of a high mass tail in the black hole mass function at $z = 0$. Consistency between the optical and X-ray census of the local black hole mass function requires an upper limit to black hole masses. This consistent picture also predicts that the slope of the $M_{\\rm bh}$-$\\sigma$ relation will evolve with redshift at the high mass end. Models of self-regulation that explain the co-evolution of the stellar component and nuclear black holes naturally provide such an upper limit. The combination of multi-wavelength constraints predicts the existence of UMBHs and simultaneously provides an upper limit to their masses. The typical hosts for these local UMBHs are likely the bright, central cluster galaxies in the nearby Universe. ", "introduction": "Observations of black hole demographics locally is increasingly providing a strong constraint on models that explain the assembly and growth of black holes in the Universe. The existence of a tight relation between the velocity dispersion of bulges and the mass of the central black hole has been reported by several authors (Merritt \\& Ferrarese 2001; Tremaine et al. 2002; Gebhardt et al. 2003). This correlation is tighter than that between the luminosity of the bulge and the mass of the central black hole (Magorrian et al. 1998). The physical processes that set up this correlation are not fully understood at the present time, although there are several proposed explanations that involve the regulation of star formation with black hole growth and assembly in galactic nuclei (Haehnelt, Natarajan \\& Rees 1998; Natarajan \\& Sigurdsson 1998; Silk \\& Rees 1999; Murray, Quataert and Thompson 2004; King 2005). Recent work by several authors has suggested that UMBHs\\footnote{Black holes with masses in excess of $5 \\times 10^9\\,M_{\\odot}$ are hereafter referred to as UMBHs.} ought to exist: Bernardi et al. (2006) show that the high velocity dispersion tail of the velocity distribution function of early-type galaxies constructed from the Sloan Digital Sky Survey (SDSS) had been under-estimated in earlier work suggestive of a corresponding high mass tail for the central black hole masses hosted in these nuclei. As first argued by Lauer et al. (2007a) and subequently by Bernardi et al. (2007) and Tundo et al. (2007), even when the scatter in the observed $M_{\\rm bh} - \\sigma$ correlation is taken into account it predicts fewer massive black holes compared to the $M_{\\rm bh} - L_{\\rm bulge}$ relation. While Bernardi et al. (2007) argue that this is due to the fact that the $\\sigma - L_{\\rm bulge}$ relation in currently available samples is inconsistent with the SDSS sample from which the distributions of $L_{\\rm bulge}$ or $\\sigma$ are based. From an early-type galaxy sample observed by HST, Lauer et al. (2007b) argue that the relation between $M_{\\rm bh} - L_{\\rm bulge}$ is likely the preferred one for BCGs (Brightest Cluster Galaxies) consistent with the harboring of UMBHs as evidenced by their large core sizes. The fact that the high mass end of the observed local black hole mass function is likely biased is a proposal that derives from optical data. Deriving the mass functions of accreting black holes from optical quasars in the Sloan Digital Sky Survey Data Release 3 (SDSS DR3), Vestergaard et al. (2008) also find evidence for UMBHs in the redshift range $0.3 \\leq z \\leq 5$. In this paper, we show that UMBHs exist using X-ray and bolometric AGN luminosity functions and for consistency with local observations of the BH mass density, an upper limit to their masses is required. To probe the high mass end of the BH mass function, in earlier works the AGN luminosity functions were simply extrapolated. This turns out to be inconsistent with local estimates of the BH mass function. Here we focus on the high mass end of the predicted local black hole mass function, i.e. extrapolation of the $M_{\\rm bh} - \\sigma$ relation to higher velocity dispersions and demonstrate that a self-limiting cut-off in the masses to which BHs grow at every epoch reconciles the X-ray and optical views. The outline of this paper is as follows: in Section 2, we briefly summarise the current observational census of black holes at high and low redshift including constraints from X-ray AGN. The pathways to grow UMBHs are described in Section 3. Derivation of the local black hole mass function from the X-ray luminosity functions of AGN is presented in Section 4. The arguement for the existence of an upper limit to black hole masses from various lines of evidence is presented in Section 5; the prospects for detection of this population is presented in Section 6 followed by conclusions and discussion. We adopt a cosmological model that is spatially flat with $\\Omega_{\\rm matter} = 0.3$; $H_0 = 70\\,{\\rm km~s^{-1}}/{\\rm Mpc}$. ", "conclusions": "The interplay between the evolution of BHs and the hierarchical build-up of galaxies appears as scaling relations between the masses of BHs and global properties of their hosts such as the BH mass vs. bulge velocity dispersion - the $M_{\\rm bh} - \\sigma_{\\rm bulge}$ relation and the BH mass vs. bulge luminosity $M_{\\rm bh} - L_{\\rm Bulge}$ relation. The low BH mass end of this relation has recently been probed by Ferrarese et al. (2006) in an ACS survey of the Virgo cluster galaxies. They find that galaxies brighter than $M_B \\sim -20$ host a supermassive central BH whereas fainter galaxies host a central nucleus, referred to as a central massive object (CMO). Ferrarese et al. report that a common $M_{\\rm CMO} - M_{\\rm gal}$ relation leads smoothly down from the scaling relations observed for more more massive galaxies. Extrapolating observed scaling relations to higher BH masses to the UMBH range, we predict that these are likely hosted by the massive, high luminosity, central galaxies in clusters with large velocity dispersions. The velocity dispersion function of early-type galaxies measured from the SDSS points to the existence of a high velocity dispersion tail with $\\sigma > 350\\,{\\rm kms}^{-1}$ (Bernardi et al. 2006). If the observed scaling relations extend to the higher mass end as well, these early-types are the most likely hosts for UMBHs. Recent simulation work that follows the merger history of cluster scale dark matter halos and the growth of BHs hosted in them by Yoo et al.(2007) also predict the existence of a rare population of local UMBHs. However, theoretical arguments suggest that there may be an upper limit to the mass of a BH that can grow in a given galactic nucleus hosted in a dark matter halo of a given spin. Clearly the issue of the existence of UMBHs is intimately linked to the efficiency of galaxy formation and the formation of the largest, most luminous and massive galaxies in the Universe. Possible explanations for the tight correlation observed between the velocity dispersion of the spheroid and black hole mass involve a range of self-regulated feedback prescriptions. An estimate of the upper limits on the black hole mass that can assemble in the most massive spheroids can be derived for all these models and they all point to the existence of UMBHs. In this paper, we have argued that while rare UMBHs likely exist, there is nevertheless an upper limit of $\\sim 10^{10}\\,\\msun$ for the mass of BHs that inhabit galactic nuclei in the Universe. We first show that our current understanding of the accretion history and mass build up of black holes allows and implies the existence of UMBHs locally. This is primarily driven by new work that predicts the formation of massive black hole seeds at high redshift (Lodato \\& Natarajan 2007) and their subsequent evolution (Volonteri, Lodato \\& Natarajan 2008). Starting with massive seeds and following their build-up through hierarchical merging in the context of structure formation in a cold dark matter dominated Universe, we show that a viable pathway to the formation of UMBHs exists. There is also compelling evidence from the observed evolution of X-ray AGN for the existence of a local UMBH population. Convolving the observed X-ray LF's of AGN, with a simple accretion model, the mass function of black holes at $z = 0$ is estimated. Mimic-ing the effect of self-regulation processes that impose an upper limit to BH masses and incorporating this into the X-ray AGN LF we find that the observed UMBH mass function at $z = 0$ is reproduced. This self-regulation limited growth is implemented by steepening the high luminosity end of the AGN LF at the bright end. We estimate the abundance of UMBHs to be $\\sim \\,7 \\times 10^{-7}\\, Mpc^{-3}$ at $z = 0$. The key prediction of our model is that the slope of the $M_{\\rm bh} - \\sigma$ relation likely evolves with redshift at the high mass end. Probing this is observationally challenging at the present time but there are several bright, massive early-type galaxies that are promising host candidates from the SDSS survey as well as a survey of bright central galaxies of nearby clusters. Observational detection of UMBHs will provide key insights into the physics of galaxy formation and black hole assembly in the Universe." }, "0808/0808.2214_arXiv.txt": { "abstract": "We present an analysis of XMM-Newton and RXTE data from three observations of the neutron star LMXB 4U~1636-536. The X-ray spectra show clear evidence of a broad, asymmetric iron emission line extending over the energy range 4--9 keV. The line profile is consistent with relativistically broadened Fe K-$\\alpha$ emission from the inner accretion disk. The Fe K-$\\alpha$ line in 4U~1636-536 is considerably broader than the asymmetric iron lines recently found in other neutron star LMXBs, which indicates a high disk inclination. We find evidence that the broad iron line feature is a combination of several K-$\\alpha$ lines from iron in different ionization states. ", "introduction": "Relativistically broadened, asymmetric Fe~K-$\\alpha$ lines from the inner accretion disk have been observed in many supermassive and stellar-mass black holes \\citep[e.g.][]{2006AN....327..943F}. In neutron star binaries, however, the iron lines are weaker, and until recently observations did not clearly reveal a relativistic line profile. \\citet{2007ApJ...664L.103B} found an asymmetric Fe~K-$\\alpha$ line in {\\it XMM-Newton} spectra of the low-mass X-ray binary (LMXB) Serpens X-1 and showed that the line profile is consistent with fluorescent iron line emission from the inner accretion disk. Similar asymmetric Fe~K-$\\alpha$ lines were found by \\citet{2008ApJ...674..415C} in {\\it Suzaku} spectra of the neutron star LMXBs Serpens X-1, 4U~1820-30, and GX~349+2. In this paper we present {\\it XMM-Newton} and {\\it RXTE} observations of the neutron star LMXB 4U~1636-536. We analyze the X-ray spectra to determine the profile of the relativistic Fe~K-$\\alpha$ line and constrain the properties of the inner accretion disk. 4U~1636-536 (V801~Ara) is a well-studied, bursting LMXB consisting of a neutron star in a 3.8-hr orbit with a 0.4 solar mass, 18th magnitude star \\citep{1990A&A...234..181V} and is located at a distance of $\\sim$6~kpc \\citep{2006ApJ...639.1033G}. The X-ray timing properties of the binary have been studied extensively. \\citet{1996ApJ...469L..17Z} and \\citet{1997ApJ...479L.141W} discovered quasi-periodic oscillations (QPOs) at kHz frequencies. The source also exhibits highly coherent burst oscillations at 581~Hz which are likely related to the rotation of the neutron star \\citep{1997IAUC.6541....1Z,2002ApJ...577..337S}. \\citet{1999ApJ...514L..31K} found that the soft X-ray emission, modulated at the kHz QPO frequency, lags behind the hard X-ray emission. A possible explanation for this phase lag is the reprocessing of hard X-rays in a cooler Comptonizing corona with a size of at most a few kilometers. \\citet{2007MNRAS.376.1139B} interpreted observations showing a decline of the QPO coherence and rms amplitude at high QPO frequencies as evidence that the inner radius of the accretion disk in 4U~1636-536 is usually larger than but sometimes approaches the innermost stable circular orbit (ISCO). However, \\citet{2006MNRAS.371.1925M} argued that this decline is not caused by effects related to the ISCO. In this paper we report on three X-ray observations of 4U~1636-536 that were carried out simultaneously with {\\it XMM-Newton} and {\\it RXTE}. We present an analysis of the X-ray spectrum over the 0.5--100~keV energy range and our results from modeling the continuum and relativistic Fe~K-$\\alpha$ line emission. We use the Fe~K-$\\alpha$ line profile to derive constraints on the disk inclination and inner disk radius. We show that the line profile is not consistent with a single relativistic Fe~K-$\\alpha$ line from a neutron star accretion disk and that at least two lines from different ionization states of iron are needed to adequately describe the line profile. Finally, we discuss the implications of our findings for measurements of neutron star radii based on Fe~K-$\\alpha$ line profiles. ", "conclusions": "We have analyzed X-ray spectra of the neutron star LMXB 4U~1636-536 obtained with {\\it XMM-Newton} and {\\it RXTE} in 2005, 2007, and 2008. The very high signal-to-noise ratio of the spectra allowed us to clearly detect a broad, relativistic Fe~K-$\\alpha$ line from the inner accretion disk. The line is significantly broader than the asymmetric Fe~K-$\\alpha$ lines recently found in other neutron star LMXBs \\citep{2007ApJ...664L.103B,2008ApJ...674..415C}. The broader line profile is likely the result of a high disk inclination in 4U~1636-536. The inclination angles derived from iron line profiles in other neutron star LMXBs have so far been comparatively low. As pointed out by \\citet{2008ApJ...674..415C}, this is likely a selection effect because the narrower lines in low inclination systems are more easily detectable. With the high signal-to-noise ratio of the 4U~1636-536 spectra, we have now been able to measure the broader line profile in a high inclination LMXB. Our analysis of the Fe~K-$\\alpha$ line profile places a lower limit of $64^\\circ$ on the disk inclination in 4U~1636-536. This limit is consistent with the 36--74$^\\circ$ constraint on the orbital inclination by \\citet{2006MNRAS.373.1235C} and with the non-detection of eclipses. When fitting the iron line profile with a single relativistic line component, we find a significant difference in the rest-frame line energy between the three observations. This difference is likely caused by a change in the ionization profile of the disk related to the change in X-ray luminosity between the three observations. The line energy derived for the second observation is close to 6.97~keV, the K-$\\alpha$ line energy of Fe~XXVI, which suggests that Fe~XXVI contributes significantly to the observed line profile. According to our fit with a disk-blackbody model, the highest temperature in the disk is $\\sim$0.9~keV. Because plasma in thermal equilibrium at this temperature does not contain a significant fraction of Fe~XXVI, is it evident that the ionization profile in the disk is strongly affected by photoionization. We find that the Fe~K-$\\alpha$ line profile for two of the observations is too broad to be adequately described by a single relativistic emission line with physically reasonable values of disk inclination and inner disk radius. The broader than expected line profile can be explained by overlapping K-$\\alpha$ lines from iron in different ionization states. The presence of multiple ionization states is also indicated by the difference in the fitted line energy between the three observations. It is evident that multiple line components need to be considered to adequately model the relativistic iron line profiles in neutron star LMXBs. We obtained reasonable line parameters when fitting two iron lines with different rest-frame energies. This is obviously an oversimplification, since many ionization states are probably contributing to the iron line profile. However, because the lines are broad and overlap considerably, it is not possible to constrain the parameters of all line components independently. Even with only two line components, the fit parameters are already strongly correlated, leading to larger parameter uncertainties than a fit with a single line component. In order to adequately model the contribution from multiple ionization states when fitting relativistic Fe~K-$\\alpha$ line profiles, improved models are needed that can predict the ionization profile in the accretion disk. The Fe~K-$\\alpha$ line profiles in neutron star LMXBs can in principle be used to place upper limits on the neutron star radius by constraining the inner disk radius \\citep{2008ApJ...674..415C}. Previously, these line profiles have been fitted with only a single line component. However, if more than one ionization state contributes significantly to the Fe~K-$\\alpha$ emission, the line profile will be broader than for a single relativistic line, and a fit with a single line model may underestimate the inner disk radius and thus the limit on the neutron star radius. This can be clearly seen for two of our observations for which the upper limit on $R_{in}$ increases from $6.3R_g$ to $11.9R_g$ and $10.9R_g$, respectively, when the iron line profile is fitted with two line components instead of a single line component. In contrast, the upper limit on $R_{in}$ for the second observation decreases from $13.3R_g$ to $9.8R_g$. It is evident that the constraints on the inner disk radius strongly depend on the assumptions made about the contributing ionization states of iron and that a better understanding of the ionization profile in the disk is needed to obtain reliable limits on the neutron star radius. The contribution of multiple ionization states may also be important for the interpretation of some iron line profiles in accreting black holes. We note that the disk inclination and inner disk radius are strongly anti-correlated when fitting broad iron line profiles, which can lead to large uncertainties of the two parameters. Prior knowledge of the disk inclination can significantly reduce the uncertainty of the inner disk radius. It was shown by \\citet{2007ApJ...656.1056L} that broad iron line features can also be produced by Compton scattering of line photons in a strong outflow. The expected line profiles for this process are generally characterized by a narrow line at $\\sim$6.4--6.6~keV from fluorescence in the outflow and a broad, redshifted component below 7~keV from downscattering of the line photons. These line profiles differ qualitatively from those found in 4U~1636-536 which show significant emission above 7~keV and no narrow line component (Figure \\ref{spec2}). We also note that Compton scattering in an outflow only contributes significantly to the iron line emission if the optical depth is at least of order unity, which requires a mass outflow rate on the order of the Eddington mass accretion rate. The X-ray luminosity we observed in 4U~1636-536 was only $\\sim$5\\% of the Eddington luminosity, which suggests that the rate of any outflow in 4U~1636-536 was significantly below the Eddington mass accretion rate. It therefore seems unlikely that a large fraction of the observed iron line emission was produced in an outflow." }, "0808/0808.3815_arXiv.txt": { "abstract": "% NGC\\,2362 is a richly populated Galactic cluster, devoid of natal molecular gas and dust. The cluster represents the final product of the star forming process and hosts an unobscured and near-complete initial mass function. NGC\\,2362 is dominated by the O9 Ib multiple star, $\\tau$ CMa, as well as several dozen unevolved B-type stars. Distributed throughout the cluster are several hundred suspected intermediate and low-mass pre-main sequence members. Various post-main sequence evolutionary models have been used to infer an age of $\\sim$5 Myr for the one evolved member, $\\tau$ CMa. These estimates are in close agreement with the ages derived by fitting pre-main sequence isochrones to the contracting, low-mass stellar population of the cluster. The extremely narrow sequence of stars, which extends more than 9 mag in the optical color-magnitude diagram, suggests that star formation within the cluster occurred rapidly and coevally across the full mass spectrum. Ground-based near infrared and H$\\alpha$ emission surveys of NGC\\,2362 concluded that most ($\\sim$90\\%) of the low-mass members have already dissipated their optically-thick, inner ($\\ll$1 AU) circumstellar disks. {\\it Spitzer} IRAC observations of the cluster have confirmed these results, placing an upper limit on the primordial, optically thick disk fraction of the cluster at $\\sim$7$\\pm$2\\%. The presence of circumstellar disks among candidate members of NGC\\,2362 is also strongly mass-dependent, such that no stars more massive than $\\sim$1.2 M$_{\\odot}$ exhibit significant infrared excess shortward of 8 $\\mu$m. NGC\\,2362 will likely remain a favored target of ground-based and space-based observations. Its well-defined upper main sequence, large population of low-mass, pre-main sequence stars, and the narrow age spread evident in the color-magnitude diagram ensure its role as a standard model of cluster as well as stellar evolution. ", "introduction": "The young cluster NGC\\,2362 in Canis Majoris (CMa) is dominated by the 4$^{th}$ mag O9 Ib multiple star, $\\tau$ CMa and several dozen B-type stars (McSwain \\& Gies 2005), spherically distributed within a volume $\\sim$3 pc in radius. Shown in Figure~1 is a 15\\arcmin$\\times$15\\arcmin\\ Second Palomar Observatory Sky Survey (POSS-II) red image of NGC\\,2362 obtained from the Digitized Sky Survey. The cluster is free of molecular gas and nebular emission and suffers very little interstellar reddening despite an accepted distance of nearly 1.5 kpc. With an age of $\\sim$5 Myr, only $\\tau$ CMa has evolved significantly away from the cluster zero-age main sequence (ZAMS). Although probably relaxed, the evaporation timescale for the cluster is significantly greater than its age, implying that few members have dispersed. In essence, an unobscured and near-complete stellar population remains around $\\tau$ CMa, making NGC\\,2362 an ideal target for initial mass function (IMF) studies (Moitinho et al. 2001). The rich history of the cluster begins with its discovery by Fr. Giovanni Battista Hodierna in Sicily during the 17$^{th}$ century using a Galilean-type refractor. His observations of nebulous objects which includes NGC\\,2362 were published in 1654. With only one member ($\\tau$ CMa) visible to the unaided eye, the cluster quickly returned to obscurity for over a century before Sir William Herschel noted its presence, entry H VII.17 in his catalog of stellar clusters and nebulae, one of the pre-cursors to Dreyer's New General Catalog (NGC). Dreyer's notes within the original NGC summarize NGC\\,2362 as a pretty large, rich cluster centered upon 30 CMa ($\\tau$). The compact nature of NGC\\,2362 is quite striking when viewed on the Palomar Observatory Sky Survey (POSS) plates. The large number of early-type stars form a luminous halo around $\\tau$ CMa, $\\sim$10\\arcmin\\ in diameter (Figure~\\ref{f1}). \\begin{figure}[h!] \\hspace{-1.0cm} \\plotfiddle{f1.ps}{11.6cm}{90.0}{58.0}{58.0}{70.0}{-15.0} \\caption[f1.ps]{A 15\\arcmin$\\times$15\\arcmin\\ Second Palomar Observatory Sky Survey (POSS-II) red image of NGC\\,2362 obtained from the Digitized Sky Survey. $\\tau$ CMa, the 4$^{th}$ mag O9 Ib star (center), is the most massive cluster member and the only star that has evolved significantly away from the ZAMS. Given the abundance of unevolved B-type members from B1 to B9, the early population of NGC\\,2362 has often been used to define the upper section of the empirically-derived ZAMS. The pre-main sequence population of NGC\\,2362 is symmetrically distributed around $\\tau$ CMa.\\label{f1}} \\end{figure} Although lacking nebulosity in the immediate cluster vicinity, just over one degree east of NGC\\,2362 extensive H~II emission is apparent on the POSS plates. This emission is part of the giant H~II region Sharpless 310, described by Sharpless (1959) as an incomplete ring 8$^{\\circ}$ in diameter. The ionizing sources for this gas are most likely 29 CMa, $\\tau$ CMa, and the early B-type members of NGC\\,2362. IRAS images of the NGC\\,2362 region reveal that the cluster is located within an evacuated cavity approximately 30\\arcmin\\ in diameter. A partial ring of dust emission, most prominent at 60~$\\mu$m, is evident to the east. The nearby dark nebula, L1667, also believed to be 1.5~kpc distant (Lada \\& Reid 1978), lies southeast of NGC\\,2362, near the variable M5 supergiant VY~CMa. Tenuous dark clouds appear to follow the contours of H~II emission on the POSS plates, perhaps remnants of the molecular cloud complex from which NGC\\,2362 formed. The OB stellar population of NGC\\,2362 may have triggered a second generation of star formation within L1660 to the northeast, which hosts the Herbig-Haro object HH~72 (Reipurth \\& Graham 1988). Shown in Figure~2 is a reproduction of a Schmidt plate from Reipurth \\& Graham (1988) centered upon L1660 and with the locations of HH~72 and three known H$\\alpha$ emission stars identified. Reipurth \\& Graham (1988) conclude that L1660 is slowly being eroded away by intense UV radiation from the OB stellar population of NGC\\,2362 and other nearby massive stars. \\begin{figure} \\centering \\includegraphics[angle=0,width=0.85\\textwidth]{f2.eps} \\caption[f2.eps]{The L1660 dark cloud and HH 72 on the SERC-J Schmidt plate from Reipurth \\& Graham (1988). The field shown is approximately 13\\arcmin$\\times$17\\arcmin. The locations of three H$\\alpha$ emission stars are also identified, one of which (labeled 1) is possibly a HAeBe star. The other two H$\\alpha$ emitters are candidate low-mass T Tauri stars. \\label{f2}} \\end{figure} ", "conclusions": "Over the last half-century, NGC\\,2362 has played an extraordinary role in our understanding of the star formation process. Early perceptions of an anomalous mass function were overturned with the advent of modern detectors. Deep CCD surveys of the cluster by Wilner \\& Lada (1991) and Moitinho et al. (2001) revealed a well-populated, pre-main sequence extending more than 9~mag to nearly the substellar limit. There are new indications, however, that a deficit of low-mass stars is present within the cluster, implying that the IMF issue has yet to be fully resolved. Recent {\\it Chandra} X-ray observations of NGC\\,2362 have added several hundred more pre-main sequence candidates to the $\\sim$100$+$ suspected members exhibiting H$\\alpha$ emission or strong Li~I $\\lambda$6708 absorption. Deeper optical and infrared surveys of the cluster will also push the source detection threshold into the brown dwarf regime, permitting a closer examination of the cluster IMF. Interest in NGC\\,2362 has also shifted to the remaining optically thick circumstellar disks around low-mass cluster members. Near infrared and H$\\alpha$ emission surveys suggest that the inner disk regions have dissipated for most ($\\sim$90\\%) of the suspected cluster members as evidenced by the decay of near infrared excess and strong H$\\alpha$ emission. {\\it Spitzer} observations are beginning to resolve the remaining questions of disk frequency within the cluster. It is perhaps somewhat ironic that in the process of identifying and characterizing the low-mass population of NGC\\,2362 over the last decade, the OB stars have been somewhat neglected. McSwain \\& Gies' (2005) recent Str\\\"omgren photometric survey of 41 OB-type stars in the cluster region found only one candidate classical Be star. A more thorough spectroscopic analysis of the B-star population is needed to confirm spectral types, evaluate membership, and to examine questions of binarity, critical to understanding the placement of stars on the ZAMS. NGC\\,2362 will likely remain a favored target for ground-based and space-based observations. Its large, statistically-significant population of low-mass, pre-main sequence stars, its well-defined upper main sequence, compact structure, and lack of circumstellar and interstellar gas and dust relative to similarly aged clusters all contribute to the cluster's unique nature. \\vspace{0.5cm} {\\bf Acknowledgments.} I wish to thank the referee for this paper, Francesco Damiani, and the editor, Bo Reipurth, for many helpful comments that significantly improved the manuscript. SED is supported by an NSF Astronomy and Astrophysics Postdoctoral Fellowship under award AST-0502381. The Digitized Sky Surveys were produced at the Space Telescope Science Institute under U.S. Government grant NAG W-2166. The images of these surveys are based on photographic data obtained using the Oschin Schmidt Telescope on Palomar Mountain and the UK Schmidt Telescope. The plates were processed into the present compressed digital form with the permission of these institutions. The Second Palomar Observatory Sky Survey (POSS-II) was made by the California Institute of Technology with funds from the National Science Foundation, the National Geographic Society, the Sloan Foundation, the Samuel Oschin Foundation, and the Eastman Kodak Corporation." }, "0808/0808.3448_arXiv.txt": { "abstract": "This paper reports on the results of a numerical investigation designed to address how the initially anisotropic appearance of a GRB remnant is modified by the character of the circumburst medium and by the possible presence of an accompanying supernova (SN). Axisymmetric hydrodynamical calculations of light, impulsive jets propagating in both uniform and inhomogeneous external media are presented, which show that the resulting dynamics of their remnants since the onset of the non-relativistic phase is different from the standard self-similar solutions. Because massive star progenitors are expected to have their close-in surroundings modified by the progenitor winds, we consider both free winds and shocked winds as possible external media for GRB remnant evolution. Abundant confirmation is provided here of the important notion that the morphology and visibility of GRB remnants are determined largely by their circumstellar environments. For this reason, their detectability is highly biased in favor of those with massive star progenitors; although, in this class of models, the beamed component may be difficult to identify because the GRB ejecta is eventually swept up by the accompanying SN. The number density of asymmetric GRB remnants in the local Universe could be, however, far larger if they expand in a tenuous interstellar medium, as expected for some short GRB progenitor models. In these sources, the late size of the observable, asymmetric remnant could extend over a wide, possibly resolvable angle and may be easier to constrain directly. ", "introduction": "\\label{int} Relativistic jets are common in the astrophysical environment. Objects known or suspected to produce them include radio galaxies and quasars \\citep{b84}, microquasars \\citep{mr99} and gamma-ray bursts \\citep{g06}. An important difference between jets of gamma-ray bursts (GRBs) and the better studied radio jets of quasars or microquasars is that active quasars often inject energy over extended periods of time into the jet while GRB sources are impulsive. Although quasar jets remain highly collimated throughout their lifetimes, GRB jets decelerate and expand significantly once they become nonrelativistic. Expansion into a uniform medium has been well studied \\citep{ap01}, but the interaction of a GRB remnant with a nonuniform medium remains poorly understood. Much of our effort in this paper is therefore dedicated to determining how the morphology and dynamics of young GRB remnants is modified by the character of the circumburst medium. Some of the questions at the forefront of attention include the effects of the external medium and the degree to which GRB remnant dynamics and structures are modified by the presence of an accompanying supernova. We address both of these issues here. Because massive stars are expected to have their close-in surroundings modified by the progenitor winds, we consider both free winds and shocked winds as possible surrounding media for the GRB remnant evolution. Detailed hydrodynamic simulations of this interaction are presented in \\S\\S \\ref{wind} and \\ref{bubble}, while a brief description of the numerical methods and the initial models is giving in \\S \\ref{nm}. For completeness, the interaction with a constant-density medium is discussed in \\S \\ref{const}. The role of supernova explosions in shaping the evolution and morphology of GRB remnants is discussed in \\S \\ref{sn}. The effects of a nonspherical circumburst medium are briefly addressed in \\S \\ref{wind}. Discussion and conclusions are presented in \\S \\ref{dis}. ", "conclusions": "" }, "0808/0808.0577_arXiv.txt": { "abstract": "{} {The stability of dissipative Taylor-Couette flows with an axial stable density stratification and a prescribed azimuthal magnetic field is considered. } { Global nonaxisymmetric solutions of the linearized MHD equations with toroidal magnetic field, axial density stratification and differential rotation are found for both insulating and conducting cylinder walls.} {Flat rotation laws such as the quasi-Kepler law are unstable against the nonaxisymmetric stratorotational instability (SRI). The influence of a current-free toroidal magnetic field depends on the magnetic Prandtl number Pm: SRI is supported by $\\rm Pm > 1$ and it is suppressed by $\\rm Pm \\lsim 1$. For too flat rotation laws a smooth transition exists to the instability which the toroidal magnetic field produces in combination with the differential rotation. This nonaxisymmetric azimuthal magnetorotational instability (AMRI) has been computed under the presence of an axial density gradient. If the magnetic field between the cylinders is not current-free then also the Tayler instability occurs and the transition from the hydrodynamic SRI to the magnetic Tayler instability proves to be rather complex. Most spectacular is the `ballooning' of the stability domain by the density stratification: already a rather small rotation stabilizes magnetic fields against the Tayler instability. An azimuthal component of the resulting electromotive force only exists for density-stratified flows. The related alpha-effect for magnetic SRI of Kepler rotation appears to be positive for negative $d\\rho/dz <0$. } {} ", "introduction": " ", "conclusions": "The stability of the dissipative Taylor-Couette flow under the joint influence of a stable vertical density stratification and an azimuthal magnetic field is considered. The problem is of interest for future laboratory experiments but also within the frame of accretion disk physics. The Kepler rotation generates strong toroidal magnetic fields dominating the poloidal components. The standard MRI which works with only axial fields may be of minor relevance for the stability of the Kepler rotation law compared with the azimuthal MRI in connection with the density stratification and Tayler instability. Mainly nonaxisymmetric `kink' modes ($m=1$) are considered but there are also examples where the axisymmetric modes are the most unstable ones. We started with a discussion of the SRI without magnetic fields. For a flat rotation law a stratification value Fr exists for which the critical Reynolds number of the rotation has a minimum (see Fig. \\ref{NN}). The SRI is basically stabilized by both too weak or too strong stratification. The instability only appears if the characteristic buoyancy time approaches the rotation period. Also if the rotation is too flat the instability disappears. The limiting ratio $\\mu_\\Omega$ strongly depends on the gap width (see Fig. \\ref{gaps}). For small gaps rotation laws with $\\mu_\\Omega>\\hat\\eta$ even prove to be unstable while for wide gaps the condition $\\mu_\\Omega<\\hat\\eta$ results for exciting SRI. Our previous finding that $\\mu_\\Omega \\lsim \\hat\\eta$ limits the SRI is reproduced for medium-sized gaps. In all cases, however, the quasi-Kepler rotation proves to be unstable. New experiments with different gap sizes and larger Reynolds numbers could verify these results. Figures~\\ref{amri} and \\ref{amri1} yield the basic results for SRI subject to toroidal fields. The magnetic field is assumed as current-free in the fluid between the cylinders, i.e. $B_\\phi \\propto 1/R$ excluding Tayler instability. Without density stratification no Rayleigh instability exists for quasi-Kepler rotation but nonaxisymmetric modes with $m=1$ are unstable as a result of the interaction of differential rotation and magnetic field (`AMRI'). The magnetic influence strongly depends on the magnetic Prandtl number. The instability needs higher Reynolds number for ${\\rm Pm}\\lsim 1$ and it needs lower Reynolds number for ${\\rm Pm}\\gsim 1$. After our experiences with MHD instabilities this is not a surprise. It is a surprise, however, that always the magnetic influence is only weak. Up to Hartmann numbers of ${\\rm Ha}\\simeq 100$ only a magnetic-induced factor of two plays a role\\footnote{${\\rm Ha}=100$ for gallium corresponds to $B\\approx 2200/R_0\\ [{\\rm Gauss}]$ with $R_0$ in cm}. Hence, our conclusion is that the SRI survives for rather high magnetic fields. For large Pm it is even supported by the toroidal magnetic field. The combination of rotation, density stratification and magnetic field leads to complex results. However, a basic observation is that the critical Reynolds numbers, if existing, above which the flow becomes unstable without a magnetic field, are increased by the stable density stratification. In contrast, the critical Hartmann numbers, if existing by the Tayler instability, above which the field becomes unstable without a rotation, do not depend on the stratification. Different transition are possible between the two limits. Generally speaking, the density stratification `balloons' the stability region and in this sense it stabilizes the flow. For slow rotation the maximal stable magnetic field exceeds the critical magnetic field without rotation while for faster rotation the maximal stable magnetic field is smaller than this critical value. Even rather slow values of the Reynolds numbers lead to a stabilization of those fields which are unstable for $\\rm Re=0$. The effect is strong for $\\rm Pm=1$ but it becomes smaller for decreasing magnetic Prandtl numbers. For steep radial profiles of the magnetic field (i.e. strong axial currents) and magnetic Prandtl numbers $\\rm Pm \\gsim 1$ one also finds the magnetic field destabilizing the Rayleigh instability, i.e. the critical Reynolds numbers with magnetic field are lower than the Reynolds numbers without magnetic field. This magnetic destabilization only exists for not too small magnetic Prandtl numbers. It exists for both uniform and density-stratified fluids (Fig. \\ref{gen}). Finally, the $\\phi$-component of the electromotive force representing the $\\alpha$-effect of the mean-field dynamo theory has been computed. The computations require an extreme degree of accuracy. The results demonstrate the importance of the density stratification for the existence of the $\\alpha$-effect. Without density stratification the correlations are vanishing in the radial average. With an axial density stratification the calculations model the polar region of a rotating sphere or disk. This interpretation accepted we found the $\\alpha_{\\phi\\phi}$ at the northern pole as {\\em positive} (Fig. \\ref{emf}). Again, the basic ingredient of this $\\alpha$-effect model is the density stratification." }, "0808/0808.2744_arXiv.txt": { "abstract": "We present 3D radiation-gasdynamical simulations of an ionization front running into a dense clump. In our setup, a B0 star irradiates an overdensity which is at a distance of $10\\,\\mathrm{pc}$ and modelled as a supercritical $100\\,\\msol$ Bonnor-Ebert sphere. The radiation from the star heats up the gas and creates a shock front that expands into the interstellar medium. The shock compresses the clump material while the ionizing radiation heats it up. The outcome of this ``cloud-crushing'' process is a fully turbulent gas in the wake of the clump. In the end, the clump entirely dissolves. We propose that this mechanism is very efficient in creating short-living supersonic turbulence in the vicinity of massive stars. \\noindent \\pacs{47.40.Nm,97.10.Bt,98.38.Am} ", "introduction": "Massive stars strongly influence the environment in which they have formed by stellar winds, ionizing radiation and supernova explosions. An ionization front which expands into the ambient interstellar medium (ISM) and hits a dense clump may compress it so heavily that gravitational collapse might be triggered. On the other hand, ionization heats up the material and can photoevaporate the clump. The remaining material of this competition may form low-mass stars~\\cite{hest05} or brown dwarfs~\\cite{whit04}. The interaction of a shock with a dense clump is called ``cloud-crushing''. The cloud-crushing scenario has been studied numerically both for supernova shocks and ionization fronts. The first extensive studies of the fate of the shocked cloud already showed that strong vortex rings can be produced~\\cite{klein94}. The mixing properties of the cloud depend sensitively on the initial density distribution~\\cite{naka06}. Furthermore, simulations of dense clumps exposed to an ionizing flux but without strong shocks show the generation of kinetic energy~\\cite{kessel03} and fragmentation of the clump~\\cite{esquivel07}. Radiation-gasdynamical simulations have also been used to match observations in H II regions, especially the Eagle Nebula~\\cite{williams01, miao06}. Although ionizing radiation injects a significant amount of energy into the ISM, it does not seem to be an important driving mechanism of interstellar turbulence on a global scale~\\cite{maclow04}. However, the cloud-crushing process generates a considerable amount of turbulence locally in the wake of the cloud. We find that the motion of the cloud material is mostly supersonic while the ambient gas behind the front moves only subsonically. The continuous heating limits the lifetime of the dense material, but the supersonic motions are maintained until the cloud disperses. This is contrary to the situation in jet-clump interactions, where the situation is less clear with some studies showing mostly subsonic motions~\\cite{baner07} while others claim supersonic velocity fields~\\cite{li07}. ", "conclusions": "We have seen that cloud-crushing by ionization fronts can lead to short-living supersonic turbulence. Altough only a minute fraction of the energy input is converted into kinetic energy, up to $60\\%$ of the affected gas is supersonic. While it is mainly the cold gas that is highly supersonic, it is the hot gas that moves the fastest. The bulk motion of the shock is an important contribution to the supersonic flow, since the transversal fluctuations are at best slightly supersonic." }, "0808/0808.3118_arXiv.txt": { "abstract": "We have developed a new two-dimensional hydrostatically-balanced isobaric hydrodynamic model for use in simulation of exoplanetary atmospheres. We apply this model to the infrared photosphere of the hot Jupiter HD 189733 b, for which an excellent 8-$\\mu$m light curve has been obtained. For reasonable parameter choices, the results of our model are consistent with these observations. In our simulations, strongly turbulent supersonic flow develops, with wind speeds of approximately 5~km~s$^{-1}$. This flow geometry causes chaotic variation of the temperature distribution, leading to observable variations in the light curve from one orbit to the next. ", "introduction": "Since the earliest dynamical models of strongly irradiated exoplanetary atmospheres were published \\citep{sho02,cho03}, it has been apparent that the expected large temperature gradients and the resulting high wind speeds could have such a large impact on these planets' appearance that the effects could be observable even at a distance of dozens of parsecs. As the models have grown increasingly sophisticated over the past six years, research has progressed along two distinct lines. \\citet{sho02}, \\citet{coo05}, \\citet{dob08} and \\citet{sho08} have produced three-dimensional models, in an effort to simulate the greatest possible range of relevant physical processes. Because of the high computational costs of such models, they have been run at comparatively low resolution. In contrast to this approach, \\citet{cho03}, \\citet{lan07}, \\citet{lan08}, and \\citet{lan08b} have chosen to employ two-dimensional models which can be run at higher resolutions. On highly irradiated planets, the radiative zone is believed to extend deep into the atmosphere, to a pressure depth of hundreds of bars \\citep{coo05, iro05, sho08}. The flow is therefore expected to be strongly stratified, with vertical motion comparatively unimportant. The assumption inherent in a two-dimensional model is that the motion at small scales which can be captured due to the finer grid spacing is more important than the vertical flow which a two-dimensional model must neglect. Nevertheless, a serious concern for two-dimensional models is the possible existence of crucial three-dimensional processes which do not require strong vertical motion in order to become significant. Furthermore, previous attempts to develop two-dimensional models have been hampered by questions regarding the validity of the physics involved. \\citet{lan07} employ a shallow-water model which is at best a first-order approximation to realistic atmospheric dynamics. While the barotropic-equivalent model used by \\citet{cho03}, formally very similar to the shallow-water equations, is on a firmer physical footing, there is some indication that the model becomes numerically unstable at high wind speeds \\citep{rau07}. In an attempt to achieve greater realism than is possible using a shallow-water model, \\citet{lan08} developed a model employing fully-compressible two-dimensional hydrodynamics. As we will show in \\S2, however, such models produce features which necessarily violate hydrostatic balance. In this paper, then, we present a two-dimensional model which must maintain hydrostatic equilibrium. It is our hope that this new model will be able to approach the rigor of a three-dimensional model, while maintaining the adaptability and speed of a two-dimensional model. While these modeling efforts have been underway since 2002, the ability to constrain these models using observations is a more recent development. Of particular interest, \\citet{knu07} have produced a map of the planet's longitudinal temperature variation based on their \\textit{Spitzer} observations of its flux in the 8-$\\mu$m band. From the depth of the secondary eclipse, they find a hemispherically-averaged day-side brightness temperature of $1205.1 \\pm 9.3$ K, while they estimate a cooler hemispherically-averaged night-side brightness temperature of $973 \\pm 33$ K, based on the flux curve. Interestingly, the hottest part of the planet is \\emph{not} directly beneath the star: the temperature maximum is offset some $30^\\circ$ east of the substellar point. Perhaps even more interestingly, the temperature minimum is offset $30^\\circ$ \\emph{west} of the antistellar point. It is clear, then, the temperature distribution is strongly influenced by planetary winds; at first glance it would appear that the flow is eastward on the day-side, while westward on the night-side. To date, models have not reproduced this flow geometry \\citep{cho03, coo05, lan07, cho08, dob08, lan08, sho08}. Additionally, these observations do not seem to be generally applicable to hot Jupiters. \\citet{har06}, for example, observed the 24-$\\mu$m flux of the giant exoplanet $\\upsilon$ Andromedae b, finding a much larger temperature contrast between the illuminated and dark hemispheres. Additionally, they found a slight \\emph{westward} displacement of the hot spot from the substellar point, although their observations did not exclude the possibility that there was no offset at all. The mid-infrared observations of HD 179949 b by \\citet{cow07} also appear to be consistent with zero phase offset. It is not clear to what extent these discrepancies result from different dynamics on the planets themselves, and to what extent they are caused by the different pressure depths under observation. In other respects, HD 189733 b remains a fairly typical hot Jupiter; its period, however, is shorter than most, with $P=2.218573 \\pm 0.000020$ d. The transit has allowed determinations of several other parameters: $R=1.154 \\pm 0.032 R_J$, $i=85.79^\\circ \\pm 0.24^\\circ$ \\citep{bak06}, $a=0.0313 \\pm 0.0004$, and $M=1.15 \\pm 0.04 M_J$ \\citep{bou05}. The orbit is assumed to be circular. HD 189733 b has not received the attention from modelers which has been bestowed on its more famous cousin, HD 209458 b. Nevertheless, the well-resolved flux curve produced by \\citet{knu07} has made it a most interesting target for simulation, and \\citet{sho08} apply their atmospheric model to HD 189733 b, with mixed success. They recover the correct eastward phase offset for the hot spot, but are unable to produce a flow pattern which causes the observed westward offset for the cold spot from the antistellar point. Their flow patterns in the upper atmosphere are in general supersonic, with very low pressures characterized by longitudinal and latitudinal flow from the day side towards the night side, while deeper layers are dominated by a supersonic eastward jet at the equator. Interestingly, despite the high wind speeds ($|\\mathbf{v}| \\gtrsim 3$ km/s) and the resulting large wind shear, no turbulence develops in their simulations. As we shall see, this does not match the results of the two-dimensional model presented in this paper. It is possible that their relatively low horizontal resolution (144x90) combined with the finite-difference differentiation employed by the ARIES/GEOS dynamical core used in their model conspire to produce sufficient numerical dissipation to prevent the development of turbulence; it is also possible (and possibly more likely) that the development of turbulence is prevented by three-dimensional effects which our two-dimensional model is unable to capture. An analytical study to constrain the conditions under which turbulence is expected to arise would surely be profitable; however, this, and indeed, the more general question of the conditions necessary for planetary-scale turbulent flow, must remain a topic for future investigation. In any case, it is clear that the \\citet{knu07} time-series is of sufficiently high quality to provide an excellent benchmark for both existing and future simulations of exoplanetary atmospheres. The ARIES/GEOS core used in \\citet{coo05} and \\citet{sho08} has been applied to the terrestrial atmospheres of Earth and Mars \\citep{sho08}, and the equivalent barotropic formulation of \\citet{cho03} -- also used in \\citet{rau07} -- has enjoyed some success in reproducing the primary features of Jupiter's dynamics. However, the conditions on many extrasolar planets are so utterly unlike anything seen in our solar system that it is advisable to include data from exoplanet observations when testing the models. This paper is organized as follows: In \\S 2, we derive in some detail the hydrodynamical core of our model, as well as providing a treatment of the radiative forcing scheme. In \\S 3, we apply our model to the atmosphere of HD 189733 b, comparing our results to those of \\citet{sho08} and to the data obtained by \\citet{knu07}. We conclude in \\S4. ", "conclusions": "While this model has not yet provided an optimized fit to the light-curve of HD 189733 b obtained by \\citet{knu07}, it does provide a reasonable explanation for a dynamical configuration which could give rise to the observed light curve. In general, other models have been able to reproduce the observed phase offset of the flux maximum, corresponding to an eastward shift of the hottest temperatures from the substellar point \\citep{coo05, sho08}. It has been more difficult to account for the apparent westward shift of the coldest temperatures from the anti-stellar point. However, the model proposed in this paper offers a physically reasonable mechanism by which the observed light curve may arise: at low pressures $p<200$ mbar, the flow tends to run from the substellar point to the antistellar point at supersonic speeds. This results in a collision between eastward winds and westward winds, which produces chaotic, turbulent flow on a large scale. As a result, there is no steady-state temperature distribution; the enormous winds cause unpredictable shifts in the temperature distribution which, in turn, alter the position of the flux minima and maxima from orbit to orbit. The phase offset of the flux minimum observed by \\citet{knu07} is not inconsistent with these results. Furthermore, this hypothesis is readily testable: turbulence on the scale predicted by our model is expected to induce shifts of several hours or more in the timing of the flux minima and flux maxima; the light-curve obtained by \\citet{knu07} indicates that it is possible to measure the timing of these flux extrema with an error of $\\lesssim \\pm 1$ hour. Therefore, if subsequent observations of HD 189733 b showed significant variations in the infrared light curve from one orbit to the next, the presence of significant large-scale turbulence would be strongly supported. Conversely, the absence of such variations would imply that turbulence on a scale necessary to explain the \\citeauthor{knu07} results is suppressed by other factors for which this simple two-dimensional model is unable to account. Some support for the model presented here may be found from the recent 24-$\\mu$m phase curve obtained by \\citet{knu08}. These observations were taken over the same portion of the orbit as the earlier 8-$\\mu$m observations \\citep{knu07}. In both cases, the flux maximum precedes secondary eclipse. In contrast to the 8-$mu$m data, however, the 24-$\\mu$m curve increases monotonically throughout the observing window -- there is no flux minimum following the transit. It is unclear whether this is due to a qualitatively different flow at the 24-$\\mu$m photosphere, or due to the type of turbulent variation suggested by the model presented in this paper. It would be useful to obtain more observations in both the 8- and 24-$\\mu$m bands so that a comparison can be made between flows on multiple orbits, but at the same atmospheric depth. In any case, it is quite clear that extensive further observations are necessary to obtain a detailed characterization of the atmospheric behavior of HD 189733 b. The model presented here represents a significant forward step from previous two-dimensional models. Although the number of free parameters in our treatment of the radiative forcing mitigates our success in fitting the \\citet{knu07} light curve, the quality of the fit is sufficient to conclude that the dynamics produced by our model offer a reasonable explanation for the observations, while a poor fit would imply that the simulated flows are excluded by the data. We can therefore say with some confidence that our model is not excluded by the observations currently available. It is also the first model -- in either two or three dimensions -- to provide an explanation for the unexpected phase offset of the flux minimum in the HD 189733 b 8-$\\mu$m curve. Despite these encouraging results, much room for improvement exists. A more sophisticated treatment of the radiative forcing is necessary. However, true radiative transfer is difficult to approximate in a two-dimensional model; among other issues, the depth to which incident radiation penetrates depends rather strongly on the wavelength. It is therefore likely that improvements in this area will require a shift to a fully three-dimensional code. Furthermore, a three-dimensional treatment is necessary to ensure that departures from stratified flow do not significantly affect the results. However, the ubiquity of turbulent flow obtained in two-dimensional simulations suggests that small-scale motion is non-negligible, so that the low horizontal resolutions seen current three-dimensional models are less than ideal. An immediate goal is therefore the production of a high-resolution three-dimensional model, including a realistic multi-wavelength treatment of radiative transfer. With the recent decision to fund a non-cryogenic \\textit{Spitzer} mission motivated at least in part by the promise of useful observations of exoplanets, the efforts of modelers to assist in the identification of interesting targets will be critical. Whatever uncertainties still plague these efforts, it is certain that the influx of \\textit{Spitzer} data over the next few years should provide rigorous tests on both existing models and on those yet to be developed." }, "0808/0808.3560_arXiv.txt": { "abstract": "{During the OGLE-2 operation, Soszynski et al.~(2003) found 3 LMC candidates for an RR Lyr-type component in an eclipsing binary system. Two of those have orbital periods that are too short to be physically plausible and hence have to be optical blends. For the third, \\targetnts, we developed a model of the binary that could host the observed RR Lyr star. After being granted HST/WFPC2 time, however, we were able to resolve 5 distinct sources within a 1.3'' region that is typical of OGLE resolution, proving that \\target is also an optical blend. Moreover, the putative eclipsing binary signature found in the OGLE data does not seem to correspond to a physically plausible system; the source is likely another background RR Lyr star. There are still no RR Lyr stars discovered so far in an eclipsing binary system.} ", "introduction": "Eclipsing binary systems (EBs) have long been recognized as one of the most astrophysically rewarding targets for the reliable determination of masses, radii, luminosities and other physical stellar properties. The precision of the derived parameters is typically better than a few percent; it enables the studies of stellar structure and evolution, and yields reliable distances that are independent of any calibrations or empirical relations. EBs thus serve as \\emph{astrophysical laboratories} suitable for the study of individual component properties. The importance of finding intrinsic variables in EBs is thus obvious: being able to obtain the fundamental properties of such stars provides improved theoretical understanding, better calibrations, and in some cases leads to improving the cosmological distance ladder \\citep{guinan1998,fitzpatrick2003}. Cepheids, for example, have been found in galactic EBs \\citep{freyhammer2005,antipin2007} as well as in the LMC \\citep{alcock2002,lepischak2004,guinan2005}; there are more than 30 known $\\delta$-Scuti type components in EBs \\citep[see e.g.][]{dallaporta2002,rodriguez2004,christiansen2007}, and several other types such as slowly pulsating B stars \\citep{pigulski2007,pilecki2007}. To date, though, no RR Lyr type star has been found in an EB. Recently, however, an accurate parallax and relative proper motion of RR Lyr itself was obtained with the HST Fine Guidance Sensor \\citep{benedict2002} that estimated its absolute magnitude of $M = 0.61 \\pm 0.1$. The Optical Gravitational Lensing Experiment (OGLE) carried out a 4.5 square degree survey of the LMC during the second phase of operations (Jan 1997 -- Nov 2000; \\citealt{soszynski2003}). They discovered 7612 RR Lyr-type objects: 5455 fundamental mode pulsators, 1655 first-overtone, 272 second-overtone, and 230 double-mode pulsators, along with several dozen other short-period pulsating variables. Three objects in their sample exhibited a superimposed RR Lyrae type variability on an eclipsing binary light curve: \\targetnts, OGLE051822.60-691817.3, and OGLE050731.10-693010.3. These could be either the results of blending, or genuine RR Lyr stars in eclipsing binaries. In the latter case it would be possible, with additional spectroscopic data, to determine a reliable mass and radius estimate of an RR Lyr star for the first time. Given their (nearly) constant mean luminosities ($\\langle L\\rangle \\sim 45L_\\odot$ for RR Lyr type ab) and easily recognizable light curves, RR Lyr stars are important for the studies of the formation and evolution of population II stars, and for determining the galactic distance scale (globular clusters, LMC, and the local group; see \\citealt{brown2004,sarajedini2006}). Although considerable effort went into calibrating the luminosity function for RR Lyr stars, it is compromised by dependences on metallicity, reddening, and possibly on period and/or pulsational modes \\citep{sollima2006}. Of the three candidates, \\citet{soszynski2003} concluded that, based on the period of the binary, the latter two are likely to be blends, leaving \\target as the remaining candidate for an EB system hosting an RR Lyr component. We conducted a thorough feasibility study for all three candidates confirming their preliminary results: only the \\target light curve could be plausibly attributed to an RR Lyr component in the EB system. The OGLE $I_C$ data (archive ID \\verb|LMC_SC6_I_433519|) were passed through the period analysis algorithm PDM \\citep{stellingwerf1978} and two distinct minima were detected: one pertaining to the RR Lyr ($P_\\mathrm{RRLyr} = 0.564876$-d) and the other to the putative EB ($P_\\mathrm{EB} = 8.92371$-d). The disentangling of the two contributions was done by \\emph{polyfit} \\citep{prsa2008} to the data phased at the $P_\\mathrm{RRLyr}$ period (cf.~Fig.~\\ref{lcs}, left). The residuals, folded at $P_\\mathrm{EB}$, exhibit an EB-like shape (Fig.~\\ref{lcs}, right). Assuming canonical values of parameters for the RR Lyr component: $\\mathscr M_\\mathrm{RRLyr} = 0.53 \\mathscr M_\\odot$ \\citep{sandage2004} and $\\mathscr R_\\mathrm{RRLyr} = 5.4 \\mathscr R_\\odot$ \\citep{marconi2005}, we were able to derive a solution consistent with the observed OGLE color of the system ($(B-V)_0 = 0.27$) and the expected absolute magnitude of an RR Lyr star ($M_V \\sim 0.5$). To solve the light curve we used {\\tt PHOEBE} \\citep{prsa2005}, an analysis suite based on the Wilson-Devinney algorithm \\citep{wd1971}. Our preliminary results indicated that the observed light curve could be plausibly explained by an EB hosting a horizontal branch (HB) primary and an RR Lyr secondary. Table \\ref{params} lists the preliminary parameters obtained, which were used to compute the model light curve depicted in Fig.~\\ref{lcs}. \\begin{figure*} \\centering \\includegraphics[height=\\textwidth,angle=-90]{figs/oglephot.eps} \\caption{Left: light curve of \\target phased with the RR Lyr pulsation period ($P_\\mathrm{RRLyr} = 0.564876$-d). The solid line represents a 2nd order polynomial chain fit obtained by \\emph{polyfit} \\citep{prsa2008}. The apparent scatter under the curve is due to the putative EB signature. Right: residual light curve after subtracting the theoretical RR Lyr pulsation fit, phased at the EB period ($P_\\mathrm{EB} = 8.92371$-d). The solid line depicts a model solution derived by {\\tt PHOEBE}.} \\label{lcs} \\end{figure*} \\begin{table} \\caption{Preliminary parameters of \\target HB--RR Lyr EB obtained by fitting the residual light curve depicted in Fig.~\\ref{lcs} (right). Values denoted with an asterisk (${}^*$) are assumed. The errors in the table are \\emph{formal}, derived from the covariance matrix.} \\label{params} \\centering \\begin{tabular}{lccc} \\hline\\hline Parameter: & & System & \\\\ & Primary (HB) & & Secondary (RRLyr) \\\\ \\hline $a$ [$R_\\odot$] & & 18.5${}^\\dagger$ & \\\\ $i$ [${}^\\circ$] & & $69.0 \\pm 1.45$ & \\\\ $T_\\mathrm{eff}$ [K] & $8300 \\pm 113$ & & $6700 \\pm 66$ \\\\ $\\mathscr M$ [$\\mathscr M_\\odot$] & 0.53${}^*$ & & 0.53${}^*$ \\\\ $\\mathscr R$ [$\\mathscr R_\\odot$] & $4.12 \\pm 0.15$ & & $5.41 \\pm 0.24$ \\\\ $M_\\mathrm{bol}$ & $0.14 \\pm 0.04$ & & $0.5^*$ \\\\ $L (I_C)$ [] & $6.36 \\pm 0.09$ & & $6.44 \\pm 0.08$ \\\\ \\hline \\multicolumn{4}{l}{\\rule{0pt}{2.6ex} ${}^\\dagger$derived from $\\mathscr M_1 + \\mathscr M_2 = 1.06 \\mathscr M_\\odot$ and period $P=8.924$-d.} \\end{tabular} \\end{table} ", "conclusions": "This paper establishes the nature of \\target as a \\emph{single} RR Lyr type star. Moreover, it shows the two-fold danger of third light contamination: 1) the binary model based on OGLE data alone provided a perfectly plausible physical description with virtually no means of resolving it without additional observations, and 2) although the residuals appeared to clearly point to the binary star signature, it seems that, somewhat ironically, the source of this secular variation is a distinct short period RR Lyr type c star. Being aware of these traps is all the more important in the era of fully automatic surveys and missions, where human supervision and a detailed object-by-object study is no longer possible. False positives present a serious challenge not only for ground-based observations, but for the upcoming space missions like Kepler and Gaia as well, emphasizing the importance of follow-up observations." }, "0808/0808.1279_arXiv.txt": { "abstract": "Depending on the density reached in the cores of neutron stars, such objects may contain stable phases of novel matter found nowhere else in the Universe. This article gives a brief overview of these phases of matter and discusses astrophysical constraints on the high-density equation of state associated with ultra-dense nuclear matter. ", "introduction": "A forefront area of modern research concerns the exploration of the properties of ultra-dense nuclear matter and the determination of the equation of state (EoS)--the relation between pressure, temperature and density--of such matter. Experimentally, relativistic heavy-ion collision experiments enable physicists to cast a brief glance at hot and ultra-dense matter for times as short as about $10^{-22}$~seconds. This is different for neutron stars, which are observed with radio and X-ray telescopes as radio pulsars and X-ray pulsars. The matter in the cores of such objects is compressed permanently to densities that may be more ten times higher than the densities inside atomic nuclei, which make neutron stars natural astrophysical laboratories that allow for a wide range of (astro) physical studies and astrophysical phenomena (Fig.\\ \\ref{eos-blaschke-fig:multifaceted2}) linked to the properties of ultra-dense nuclear matter and its associated EoS.\\cite{glen97:book,Weber:1999qn,blaschke01:trento,% weber05:ppnp,Sedrakian:2006mq,page06:review,weber07:erice,Lattimer07:a,% Schaffner07:a} \\begin{figure}[tb] \\begin{center} \\includegraphics[keepaspectratio,width=0.65\\textwidth,angle=0] {multifaceted3-eos-blaschke.eps} \\caption[]{The multifaceted connection between high-density nuclear matter and neutron (compact) star phenomena.\\cite{weber05:ppnp}} \\label{eos-blaschke-fig:multifaceted2} \\end{center} \\end{figure} Of particular interest are neutron stars whose observed properties deviate significantly from the norm. Examples of such neutron stars are PSR J0751+1807 whose mass is $2.1 \\pm 0.2~M_\\odot$,\\cite{NiSp05} neutron star RX J1856.5-3754 whose radius may be $\\gsimB 13$~km,\\cite{Trumper:2003we} and XTE J1739-285 whose rotation period may be as small as 0.89~ms.\\cite{Kaaret:2006gr} As discussed in this paper, such neutron star data provide an excellent opportunity to gain profound insight into the properties of nuclear matter at most extreme conditions of density.\\cite{Klahn:2006ir,Klahn:2006iw} ", "conclusions": "" }, "0808/0808.1423_arXiv.txt": { "abstract": "We present in this paper the first results of a spectropolarimetric analysis of a small sample ($\\sim20$) of active stars ranging from spectral type M0 to M8, which are either fully-convective or possess a very small radiative core. This study aims at providing new constraints on dynamo processes in fully-convective stars. The present paper focuses on 5 stars of spectral type $\\sim$M4, i.e. with masses close to the full convection threshold ($\\simeq 0.35~\\msun$), and with short rotational periods. Tomographic imaging techniques allow us to reconstruct the surface magnetic topologies from the rotationally modulated time-series of circularly polarised profiles. We find that all stars host mainly axisymmetric large-scale poloidal fields. Three stars were observed at two different epochs separated by $\\sim$1~yr; we find the magnetic topologies to be globally stable on this timescale. We also provide an accurate estimation of the rotational period of all stars, thus allowing us to start studying how rotation impacts the large-scale magnetic field. ", "introduction": "\\label{sec:intro} Magnetic fields play a key role in every phase of the life of stars and are linked to most of their manifestations of activity. Since \\cite{Larmor19} first proposed that electromagnetic induction might be the origin of the Sun's magnetic field, dynamo generation of magnetic fields in the Sun and other cool stars has been a subject of constant interest. The paradigm of the $\\alpha\\Omega$ dynamo, i.e. the generation of a large-scale magnetic field through the combined action of differential rotation ($\\Omega$ effect) and cyclonic convection ($\\alpha$ effect), was first proposed by \\cite{Parker55} and then thoroughly debated and improved \\citep[e.g.,][]{Babcock61, Leighton69}. A decade ago, helioseismology provided the first measurements of the internal differential rotation in the Sun and thus revealed a thin zone of strong shear at the interface between the radiative core and the convective envelope. During the past few years, theoreticians pointed out the crucial role for dynamo processes of this interface -- called the tachocline -- being the place where the $\\Omega$ effect can amplify magnetic fields (see \\citealt{Charbonneau05} for a review of solar dynamo models). Among cool stars, those with masses lower than about 0.35~\\msun\\ are fully-convective \\citep[e.g.,][]{Chabrier97}, and therefore do not possess a tachocline; some observations further suggest that they rotate almost as rigid bodies \\citep{Barnes05}. However, many fully-convective stars are known to show various signs of activity such as radio, Balmer line, and X-ray emissions \\citep[e.g.,][]{Joy49, Lovell63, Delfosse98, Mohanty03, West04}. Magnetic fields have been directly detected thanks to Zeeman effect on spectral lines, either in unpolarised light \\citep[e.g.,][]{Saar85, Johns96, Reiners06}, or in circularly polarised spectra \\citep[][]{Donati06}. The lack of a tachocline in very-low-mass stars led theoreticians to propose non-solar dynamo mechanism in which cyclonic convection and turbulence play the main roles while differential rotation only has minor effects \\citep[e.g.,][]{Durney93}. During past few years, several semi-analytical approaches and MHD simulations were developed in order to model the generation of magnetic fields in fully-convective stars. Although they all conclude that fully-convective stars should be able to produce a large-scale magnetic field, they disagree on the properties of such a field, and the precise mechanisms involved in the dynamo effect remain unclear. Mean-field modellings by \\cite{Kuker05} and \\cite{Chabrier06} assumed solid-body rotation and found $\\alpha^2$ dynamo generating purely non-axisymmetric large-scale fields. Subsequent direct numerical simulations diagnose either ``antisolar'' differential rotation (i.e. poles faster than the equator) associated with a net axisymmetric poloidal field \\citep[e.g.,][]{Dobler06}; or strongly quenched ``solar'' differential rotation (i.e. the equator faster than the poles) and a strong axisymmetric toroidal field component \\citep[e.g.,][]{Browning08}. The first detailed observations of fully-convective stars do not completely agree with any of these models. Among low-mass stars, differential rotation appears to vanish with increasing convective depth \\citep{Barnes05}. This result is further confirmed by the first detailed spectropolarimetric observations of the very active fully-convective star V374~Peg by \\cite{Donati06} and \\cite{Morin08} (hereafter M08) who measure very weak differential rotation (about 1/10th of the solar surface shear). These studies also report a strong mostly axisymmetric poloidal surface magnetic field stable on a timescale of 1~yr on V374~Peg, a result which does not completely agree with any of the existing theoretical predictions. V374~Peg being a very fast rotator, observations of fully-convective stars with longer rotation periods are necessary to generalise these results. In order to provide theoretical models and numerical simulations with better constraints, it is necessary to determine the main magnetic field properties -- topology and time-variability -- of several fully-convective stars, and to find out their dependency on stellar parameters -- mass, rotation rate, and differential rotation. In this paper, we present and analyse the spectropolarimetric observations of a small sample of stars just around the limit to full convection (spectral types ranging from M3 to M4.5), collected with ESPaDOnS and NARVAL between 2006 Jan and 2008 Feb. Firstly, we briefly present our stellar sample, and our observations are described in a second part. We then provide insight on the imaging process and associated physical model. Afterwards, we present our analysis for each star of the sample. Finally, we discuss the global trends found in our sample and their implications in the understanding of dynamo processes in fully-convective stars. ", "conclusions": "\\label{sec:disc} \\begin{table*} \\caption[]{Magnetic quantities derived from our study. For each star, different observation epochs are presented separately. In columns 2--5 we report quantities from Table~\\ref{tab:sample}, respectively the stellar mass, the rotation period (with an accuracy of 2 digits), the effective Rossby number and the X-ray to bolometric luminosity ratio. Columns 6, 7 and 8 mention the Stokes $V$ filling factor, the reconstructed magnetic energy and the average magnetic flux. Columns 9--13 list the percentage of reconstructed magnetic energy respectively lying in poloidal, dipole (poloidal and $\\ell=1$), quadrupole (poloidal and $\\ell=2$), octupole (poloidal and $\\ell=3$) and axisymmetric modes ($m=0$ / $m < \\ell/2$).} \\begin{tabular}{ccccccccccccc} \\hline Name & Mass & \\Prot & $Ro$ & log$R_X$ & $f_V$ & $$ & $$ & pol. & dipole & quad. & oct. & axisymm. \\\\ & (\\msun) & (d) & ($10^{-2}$) & & & ($\\rm10^5\\,G^2$) & (kG) & (\\%) & (\\%) & (\\%) & (\\%) & (\\%) \\\\ \\hline EV~Lac (06) & 0.32 & 4.38 & 6.8 & -3.3 & 0.11 & 4.48 & 0.57 & 87 & 60 & 13 & 3 & 33/36\\\\ \\phantom{EV~Lac} (07) & -- & -- & -- & -- & 0.10 & 3.24 & 0.49 & 98 & 75 & 10 & 3 & 28/31\\\\ YZ~CMi (07) & 0.31 & 2.77 & 4.2 & -3.1 & 0.11 & 5.66 & 0.56 & 92 & 69 & 10 & 5 & 56/61\\\\ \\phantom{YZ~CMi} (08) & -- & -- & -- & -- & 0.11 & 4.75 & 0.55 & 97 & 72 & 11 & 8 & 85/86\\\\ AD~Leo (07) & 0.42 & 2.24 & 4.7 & -3.2 & 0.14 & 0.61 & 0.19 & 99 & 56 & 12 & 5 & 95/97\\\\ \\phantom{AD~Leo} (08) & -- & -- & -- & -- & 0.14 & 0.61 & 0.18 & 95 & 63 & 9 & 3 & 85/88\\\\ EQ~Peg~A (06) & 0.39 & 1.06 & 2.0 & -3.0 & 0.11 & 2.73 & 0.48 & 85 & 70 & 6 & 6 & 69/70\\\\ EQ~Peg~B (06) & 0.25 & 0.40 & 0.5 & -3.3 & na & 2.38 & 0.45 & 97 & 79 & 8 & 5 & 92/94\\\\ V374~Peg (05) & 0.28 & 0.45 & 0.6 & -3.2 & na & 6.55 & 0.78 & 96 & 72 & 12 & 7 & 75/76\\\\ \\phantom{V374~Peg} (06) & -- & -- & -- & -- & na & 4.60 & 0.64 & 96 & 70 & 17 & 4 & 76/77\\\\ \\hline \\label{tab:syn} \\end{tabular} \\end{table*} \\begin{figure*} \\center{% \\includegraphics[scale=0.60]{fig/plotMP.eps} \\caption[]{Properties of the magnetic topologies of M dwarfs as a function of rotation period and stellar mass. Larger symbols indicate larger magnetic fields while symbol shapes depict the different degrees of axisymmetry of the reconstructed magnetic field (from decagons for purely axisymmetric fields to sharp stars for purely non axisymmetric fields). Colours illustrate the field configuration (dark blue for purely toroidal fields, dark red for purely poloidal fields and intermediate colours for intermediate configurations). Solid lines represent contours of constant Rossby number $Ro=0.1$ and $0.01$ respectively corresponding approximately to the saturation and super-saturation thresholds \\citep[e.g.,][]{Pizzolato03}. The theoretical full-convection limit ($\\mstar \\simeq0.35\\msun$, \\citealt{Chabrier97}) is plotted as a horizontal dashed line.} \\label{fig:plotMP}} \\end{figure*} Spectropolarimetric observations of a small sample of active M dwarfs around spectral type M4 were carried out with ESPaDOnS at CFHT and NARVAL at TBL between 2006 Jan and 2008 Feb. Strong Zeeman signatures are detected in Stokes $V$ spectra for all the stars of the sample. Using ZDI, with a Unno-Rachkovsky's model modified by two filling factors, we can fit our Stokes $V$ time series. It can be seen on Fig.~\\ref{fig:zdi_spec_adleo}, \\ref{fig:zdi_spec_evlac}, \\ref{fig:zdi_spec_yzcmi} and \\ref{fig:zdi_spec_eqpegab} that rotational modulation is indeed mostly modelled by the imaging code. From the resulting magnetic maps, we find that the observed stars exhibit common magnetic field properties. (a) We recover mainly poloidal fields, in most stars the observations can be fitted without assuming a toroidal component. (b) Most of the energy is concentrated in the dipole modes, i.e. the lowest order modes. (c) The purely axisymmetric component of the field ($m=0$ modes) is widely dominant except in EV~Lac. These results confirm the findings of M08, i.e. that magnetic topologies of fully-convective stars considerably differ from those of warmer G and K stars which usually host a strong toroidal component in the form of azimuthal field rings roughly coaxial with the rotation axis \\citep[e.g.,][]{Donati03a}. Table~\\ref{tab:syn} gathers the main properties of the reconstructed magnetic fields and Figure~\\ref{fig:plotMP} presents them in a more visual way. We can thus suspect some trends: (a) The only partly-convective stars of the sample, AD~Leo, hosts a magnetic field with similar properties to the observed fully-convective stars. The only difference is that compared to fully-convective stars of similar $Ro$, we recover a significantly lower magnetic flux on AD~Leo, indicating that the generation of a large-scale magnetic field is more efficient in fully-convective stars. This will be confirmed in a future paper by analysing the early M stars of our sample. (b) We do not observe a growth of the reconstructed large-scale magnetic flux with decreasing Rossby number, thus suggesting that dynamo is already saturated for fully-convective stars having rotation periods lower than 5~\\d, in agreement with \\cite{Pizzolato03} and \\cite{Kiraga07}. Further confirmation from stars with $\\Prot \\gtrsim10~\\d$ is needed. This is supported by the high X-ray fluxes we report, all lying in the saturated part of the rotation-activity relation with log$R_X\\simeq-3$ \\citep[e.g.,][]{James00}. AD~Leo also exhibits a saturated X-ray luminosity despite a significantly weaker reconstructed magnetic field, indicating that the coronal heating is not directly driven by the large-scale magnetic field. (c) The only star showing strong departure from axisymmetry is EV~Lac, i.e. the slowest rotator (though lying in the saturated regime with $Ro=0.07$). Further investigation is needed to check if this a general result for fully-convective stars having $\\Prot \\gtrsim4~\\d$. The large-scale magnetic fluxes we report here range from 0.2 to 0.8~\\kG. For AD~Leo, EV~Lac and YZ~CMi, previous measurements from Zeeman broadening of atomic or molecular unpolarised line profiles report significantly higher overall magnetic fluxes (several \\kG) \\citep[e.g.,][]{Saar85, Johns96, Reiners07}. We therefore conclude that a significant part of the magnetic energy lies in small-scale fields. Even for the fast rotators EQ~Peg~A and B and V374~Peg for which ZDI is sensitive to scales corresponding to spherical harmonics up to order $\\ell=\\,$12, 20 and 25 (cf. M08), respectively, we reconstruct a large majority of the magnetic energy in modes of order $\\ell\\leq3$. This suggests that the magnetic features we miss with ZDI lie at scales corresponding to $\\ell > 25$ in the reconstructed magnetic fields of mid-M dwarfs. Three stars of the sample have been observed at two different epochs separated by about 1~yr. AD~Leo, EV~Lac, and YZ~CMi exhibit only faint variations of their magnetic topology during this time gap, the overall magnetic configuration remained stable similarly to the behaviour of V374~Peg (cf. M08). This is at odds with what is observed in more massive active stars, whose magnetic fields reportedly evolve significantly on time-scales of only a few months \\citep[e.g.,][]{Donati03a}. For three stars of our sample we are able to measure differential rotation and find that our data are compatible with solid-body rotation. In addition, for EV~Lac and YZ~CMi we infer that differential rotation is at most of the order of a few \\mrpd\\ i.e. significantly weaker than in the Sun and apparently lower than in V374~Peg (cf.~M08). This is further confirmed by the fact that the rotation periods we find are in good agreement with photometric periods previously published in the literature (whenever reliable). This result is consistent with the conclusions of the latest numerical dynamo simulations in fully convective dwarfs with $Ro\\simeq0.01$ \\citep{Browning08} showing that (i) strong magnetic fields are efficiently produced throughout the whole star (with the magnetic energy being roughly equal to the convective kinetic energy as expected from strongly helical flows, i.e., with small $Ro$) and that (ii) these magnetic fields successfully manage to quench differential rotation to less than a tenth of the solar shear (as a result of Maxwell stresses opposing the equatorward transport of angular momentum due to Reynolds stresses). However, these simulations predict that dynamo topologies of fully convective dwarfs should be mostly toroidal, in contradiction with our observations showing strongly poloidal fields in all stars of the sample; the origin of this discrepancy is not clear yet. Our study of Stokes $I$ and $V$ time-series allows to measure both the rotational period (\\Prot) and the projected equatorial velocity (\\vsini) of the sample, from which we can straightforwardly deduce the \\rsini. \\Prot\\ is well constrained by our data sets (see the error-bars in Tab.~\\ref{tab:sample}), therefore the incertitude on \\rsini\\ essentially comes from the determination of \\vsini\\ ($\\sigma\\simeq1~\\kms$). This leads to an important incertitude on the \\rsini\\ deduced for slowly rotating stars. As explained in M08, for V374~Peg we find a \\rsini\\ significantly greater than the predicted radius . Here (except for AD~Leo which is seen nearly pole-on) we find $\\rsini \\simeq \\rstar$ (cf. Tab.~\\ref{tab:sample}), suggesting radii larger than the predicted ones. This is consistent with the findings of \\cite{Ribas06} on eclipsing binaries, further confirmed on a sample of single late-K and M dwarfs by \\cite{Morales08}, that active low-mass stars exhibit significantly larger radii and cooler \\teff\\ than inactive stars of similar masses. \\cite{Chabrier07} proposed in a phenomenological approach that a strong magnetic field may inhibit convection and produce the observed trends. This back-reaction of the magnetic field on the star's internal structure may be associated with the dynamo saturation observed in our sample (see above), and with the frozen differential rotation predicted by \\cite{Browning08} when the magnetic energy reaches equipartition (with respect to the kinetic energy). We also detect significant RV variations in our sample (with peak-to-peak amplitude of up to 700~\\ms). We observe the largest RV variations on the star having the strongest large-scale magnetic field (YZ~CMi). This suggests that although the relation between magnetic field measurements and RV is not yet clear, these smooth fluctuations in RV are due to the magnetic field and the associated activity phenomena. Therefore, if we can predict the RV jitter due to a given magnetic configuration, spectropolarimetry may help in refining RV measurements of active stars, thus allowing to detect planets orbiting around M dwarfs. The study presented through this paper aims at exploring the magnetic field topologies of a small sample of very active mid-M dwarfs, i.e. stars with masses close the full-convection threshold. Forthcoming papers will extend this work to both earlier (partly-convective) and later M dwarfs, in order to provide an insight on the evolution of magnetic topologies with stellar properties (mainly mass and rotation period). We thus expect to provide new constraints and better understanding of dynamo processes in both fully and partly convective stars." }, "0808/0808.1109_arXiv.txt": { "abstract": "The efficiency of star formation governs many observable properties of the cosmological galaxy population, yet many current models of galaxy formation largely ignore the important physics of star formation and the interstellar medium (ISM). Using hydrodynamical simulations of disk galaxies that include a treatment of the molecular ISM and star formation in molecular clouds (Robertson \\& Kravtsov 2008), we study the influence of star formation efficiency and molecular hydrogen abundance on the properties of high-redshift galaxy populations. In this work, we focus on a model of low-mass, star forming galaxies at $1\\lesssim z\\lesssim2$ that may host long duration gamma-ray bursts (GRBs). Observations of GRB hosts have revealed a population of faint systems with star formation properties that often differ from Lyman-break galaxies (LBGs) and more luminous high-redshift field galaxies. Observed GRB sightlines are deficient in molecular hydrogen, but it is unclear to what degree this deficiency owes to intrinsic properties of the galaxy or the impact the GRB has on its environment. We find that hydrodynamical simulations of low-stellar mass systems at high-redshifts can reproduce the observed star formation rates and efficiencies of GRB host galaxies at redshifts $1\\lesssim z \\lesssim2$. We show that the compact structure of low-mass high-redshift GRB hosts may lead to a molecular ISM fraction of a few tenths, well above that observed in individual GRB sightlines. However, the star formation rates of observed GRB host galaxies imply molecular gas masses of $10^{8}-10^{9}~M_{\\odot}$ similar to those produced in the simulations, and may therefore imply fairly large average H$_{2}$ fractions in their ISM. ", "introduction": "To improve the physical description of star formation in hydrodynamical simulations of galaxies, \\cite[Robertson \\& Kravtsov (2008)]{robertson2008a} implemented a new model for the ISM that includes low-temperature ($T<10^{4}$K) cooling, directly ties the star formation rate to the molecular gas density, and accounts for the destruction of molecular hydrogen by an interstellar radiation field (ISRF) from young stars. They used simulations to study the relation between star formation and the ISM in galaxies and demonstrated that, for the first time, their new model simultaneously reproduces the molecular gas and total gas Kennicutt-Schmidt (KS) relations, the connection between star formation and disk rotation, and the relation between interstellar pressure and the fraction of gas in molecular form \\cite[(e.g. Wong \\& Blitz 2002, Blitz \\& Rosolowsky 2006)]{wong2002a,blitz2006a}. The capability of this model to reproduce both the star formation efficiency and molecular abundance of nearby systems makes it useful for simulating low-mass galaxies that have suppressed H$_{2}$ abundances (and whose star formation rates would be overestimated in common treatments of star formation based on the KS relation) and high-redshift galaxies whose structural properties may vary substantially from local systems (and may therefore not have the same KS relation normalization). The model should be especially useful for studying low-mass galaxies at high-redshift, such as long duration gamma-ray burst (GRB) host galaxies at $1\\lesssim z \\lesssim2$, which is the focus of this work. The highly-energetic phenomena known as GRBs were discovered over forty years ago \\cite[(Klebesadel et al. 1973)]{klebesadel1973a}, but their extragalactic origin was confirmed only in the last decade \\cite[(e.g., Metzger et al. 1997)]{metzger1997a}. Since then, the properties of the cosmological population of galaxies that host GRBs have been increasingly well-studied \\cite[(e.g., Bloom et al. 2002, Le Floc'h et al. 2006, Prochaska et al. 2006, Berger et al. 2007a,b)]{bloom2002a,le_floch2006a,prochaska2006a,berger2007a,berger2007b}. Recently, interest in long duration GRB galaxy hosts as possible tracers of the global star formation history of the universe has motivated systematic studies of their star formation efficiencies and stellar masses \\cite[(Castro Cer\\'on et al. 2008, Savaglio et al. 2008)]{castro_ceron2008a,savaglio2008a}. These studies have found that high-redshift GRB hosts have small stellar masses ($\\log M_{\\star}\\sim 9.3$) and moderate star formation rates ($\\mathrm{SFR}\\sim2.5~M_{\\odot}~\\mathrm{yr}^{-1}$). Compared with other high-redshift galaxy populations, GRB hosts tend to have lower star formation rates at fixed stellar mass compared with Lyman-break galaxies and lower stellar masses at fixed star formation rate compared with field galaxies \\cite[(for details, see Savaglio et al. 2008)]{savaglio2008a}. Spectroscopic studies of GRB sightlines have provided additional information about the post-explosion character of the host galaxy ISM. \\cite[Tumlinson et al. (2007)]{tumlinson2007a} failed to detect H$_{2}$ in five GRB sightlines and suggested that low metallicity and large far ultraviolet ISRF strengths ($10-100\\times$ the Milky Way value) were responsible for destroying molecular hydrogen in GRB hosts. They interpreted the lack of vibrationally excited H$_{2}$ lines as evidence against the GRB destroying its parent molecular cloud, but noted various caveats to this conclusion such as the parent cloud size or cloud photodissociation before to the GRB. \\cite[Whalen et al. (2008)]{whalen2008a} used one-dimensional radiative hydrodynamical calculations to show that GRBs can ionize nearby neutral hydrogen, but suggested that an additional ISRF is necessary to remove molecular hydrogen from the nearby ISM. \\cite[Prochaska et al. (2008)]{prochaska2008a} studied NV absorption in GRB sightlines, and argued that if nitrogen ionization by GRB afterglows leads to NV absorption then the observations support a scenario where dense, molecular cloud-like environments serve as the sites of GRBs. Given the increasingly detailed studies of GRB hosts, their interesting ISM and star formation properties, and their low stellar masses, a theoretical study of GRB host galaxy analogues using hydrodynamical simulations that include a treatment of the molecular ISM is warranted. Below, we present simulations of a model GRB host galaxy that include a prescription for the molecular ISM and star formation in molecular clouds \\cite[(Robertson \\& Kravtsov 2008)]{robertson2008a}. We use the simulations to examine the star formation efficiency and molecular hydrogen content of galaxies with structural properties similar to those expected for low-mass galaxies at $1\\lesssim z \\lesssim 2$. Below, we discuss our methodology and present some initial results. ", "conclusions": "" }, "0808/0808.1615_arXiv.txt": { "abstract": "Viable alternatives to astrophysical black holes include hyper-compact objects without horizon, such as gravastars, boson stars, wormholes and superspinars. The authors have recently shown that typical rapidly-spinning gravastars and boson stars develop a strong instability. That analysis is extended in this paper to a wide class of horizonless objects with approximate Kerr-like geometry. A detailed investigation of wormholes and superspinars is presented, using plausible models and mirror boundary conditions at the surface. Like gravastars and boson stars, these objects are unstable with very short instability timescales. This result strengthens previous conclusions that observed hyper-compact astrophysical objects with large rotation are likely to be black holes. ", "introduction": "Astrophysical Black Holes (BHs) are believed to be common objects in galaxies. Their mass is expected to span many orders of magnitude, from a fraction of the solar mass (primordial BHs in the galactic halo) to few solar masses (stellar BHs in the galactic plane) up to several billions of solar masses (supermassive BHs in galactic centers). Their angular momentum should be close to the extremal limit due to accretion and mergers \\cite{Gammie:2003qi, Merritt:2004gc}. For example, if quasars are powered by supermassive BHs, astrophysical observations suggest that they should be rotating near the Kerr bound \\cite{Wang:2006bz}. Unquestionable observational evidence of the existence of BHs is still lacking \\cite{Narayan:2005ie, Abramowicz:2002vt, Lasota:2006jh}. Current astrophysical data cannot rule out ``BH forgeries'', i.e.\\ hyper-compact objects with redshift and geodesics similar to those of BHs, but lacking an event horizon. Several models of hyper-compact objects with these characteristics have been known in the literature for some time. Among these models, gravastars \\cite{Chapline:2000en,Mazur:2001fv} and boson stars \\cite{bosonstars, Berti:2006qt} have been proposed as the most viable alternatives to astrophysical BHs. The authors recently showed that rapidly spinning gravastars and boson stars may develop a strong ergoregion instability \\cite{Cardoso:2007az}. Their typical instability timescales are of order of $0.1$ seconds to 1 week for objects with mass $M = 1 - 10^6 M_{\\odot}$. Therefore, observed astrophysical hyper-compact objects are likely not to be gravastars nor boson stars. The purpose of this paper is to compute the ergoregion instability for other horizonless, Kerr-like hyper-compact objects: wormholes and superspinars. Wormholes can be objects even simpler than BHs \\cite{Morris:1988cz, visserbook, Lemos:2003jb}. They are infinitesimal variations of the Schwarzschild space-time which may be indistinguishable from BHs \\cite{Damour:2007ap}. In a string theory context, the fuzzball model replaces BHs by horizonless structures \\cite{Mathur:2005zp}. The BH-like geometry emerges in a coarse-grained description which ``averages'' over horizonless geometries and produces an effective horizon at a radius where the individual microstate geometries start to differ. Superspinars are solutions of the gravitational field equations that violate the Kerr bound. These geometries could be created by high energy corrections to Einstein gravity such as those present in string-inspired models \\cite{Gimon:2007ur, Matsas:2007bj}. Superspinars are expected to have compactness of the order of extremal rotating Kerr BHs and to exist in any mass range. A rigorous analysis of the ergoregion instability for these models is a non-trivial task; known wormhole solutions are special non-vacuum solutions of the gravitational field equations. Thus their investigation requires a case-by-case analysis of the stress-energy tensor. Exact solutions of four-dimensional superspinars are not known. To overcome these difficulties, the following analysis will focus on a simple model which captures the essential features of most Kerr-like horizonless hyper-compact objects. Superspinars and rotating wormholes will be modeled by the exterior Kerr metric down to their surface, where Dirichlet boundary conditions are imposed. This problem is very similar to Press and Teukolsky's ``BH bomb'' \\cite{bhbombPress, Cardoso:2004nk}, i.e. a rotating BH surrounded by a perfectly reflecting mirror with its horizon replaced by a reflecting surface. These boundary conditions are perfect mirror conditions and require a reflection coefficient $R=1$. In a more realistic model $R<1$ and a certain transmittance $T=1-R$ should be taken into account, which will in principle decrease the strength of the ergoregion instability. We argue that the qualitative behavior of the instability is the same as long as the reflection or the superradiant amplification are large enough. Letting $\\rho$ be a superradiant factor, one expects an ergoregion instability to develop whenever $\\rho (1-T)>1$. In the perfect mirror limit $T=0$ and the superradiant condition is simply $\\rho>1$. The general case can be handled using both the analytical and numerical techniques presented here. In Section \\ref{sec:superspinars} we introduce the class of objects we will deal with in this work. They are general approximations to superspinars and wormholes with the basic key features retained. In Section \\ref{sec:wh} we show how to solve for the instability analytically in two different regimes. The details of these computations are left for Appendices \\ref{app:slow rotation} and \\ref{app:extremal}. These approximations are compared with numerical results in Section \\ref{sec:num}, where we also show that another kind of instability sets in for general naked singularities. This ``algebraic'' instability can be computed algebraically in the Kerr geometry. We close with a brief discussion of our results. ", "conclusions": "} This paper presented a general method for investigating the ergoregion instability of ultra-compact, horizonless Kerr-like objects. The essential features of these objects have been captured by a simple model whose physical properties are largely independent from the dynamical details of the gravitational system. The method has been applied to superspinars and rotating wormholes. Numerical and analytic results show that the ergoregion instability of these objects is extremely strong for any value of their angular momentum, with timescales of order $10^{-5}$ seconds for a $1 M_\\odot$ star and $10$ seconds for a $M=10^6 M_{\\odot}$ star. The above investigation confirms previous results for gravastars and boson stars \\cite{Cardoso:2007az}, namely that exotic objects without event horizon are likely to be ruled out as viable candidates for astrophysical hyper-compact objects." }, "0808/0808.0519_arXiv.txt": { "abstract": "I describe the IR and X-ray observational campaign we have undertaken for the purpose of determining the nature of the faint discrete X-ray source population discovered by Chandra in the Galactic Center (GC). Data obtained for this project includes a deep Chandra survey of the Galactic Bulge; deep, high resolution IR imaging from VLT/ISAAC, CTIO/ISPI, and the UKIDSS Galactic Plane Survey (GPS); and IR spectroscopy from VLT/ISAAC and IRTF/SpeX. By cross-correlating the GC X-ray imaging from Chandra with our IR surveys, we identify candidate counterparts to the X-ray sources via astrometry. Using a detailed IR extinction map, we are deriving magnitudes and colors for all the candidates. Having thus established a target list, we will use the multi-object IR spectrograph FLAMINGOS-2 on Gemini-South to carry out a spectroscopic survey of the candidate counterparts, to search for emission line signatures which are a hallmark of accreting binaries. By determining the nature of these X-ray sources, this FLAMINGOS-2 Galactic Center Survey will have a dramatic impact on our knowledge of the Galactic accreting binary population. ", "introduction": "The unprecedented sensitivity and angular resolution of \\Chandra has been utilized by Wang \\etal\\ (2002; hereafter W02 \\cite{wang}) and Muno \\etal\\ (2003; hereafter M03 \\cite{muno03}) to investigate the X-ray source population of the Galactic Center (GC). The W02 ACIS-I survey of the central 0.8\\degs$\\times$2\\degs of the GC revealed a population of $\\sim$800 previously undiscovered discrete weak sources with X-ray luminosities of $10^{32}-10^{35}$\\ergs. M03 imaged the central 40 pc$^{2}$ (at 8 kpc) around Sgr A*, finding an additional $\\sim$2300 discrete point sources down to a limiting flux of $10^{31}$ erg/s. More recently, a deeper \\Chandra survey of the central $\\sim$1\\degs around Sgr A* has been obtained; the combination of all of these surveys has revealed a total of $>$10,000 discrete sources (Figure 1; Muno \\etal~ 2008, hereafter M08 \\cite{muno08}). The harder ($\\geq$3 keV) X-ray sources are likely to be at the distance of the GC, while the softer sources are likely to be foreground X-ray active stars or CVs within a few kpc of the Sun. Some individual sources have been identified as X-ray transients, high-mass stars, LMXBs, and CVs. However, the nature of the majority of these newly detected sources is as yet unknown. \\begin{figure} \\includegraphics[width=0.99\\textwidth]{rmb_fig1.jpg} \\caption{ACIS-I mosaic of the GC, combining the surveys of W02, M03, and M08, as well as additional archival \\Chandra data in the region \\cite{muno08}. Red is 1-3 keV, green is 3-5 keV, and blue is 5-8 keV.} \\end{figure} ", "conclusions": "" }, "0808/0808.2807_arXiv.txt": { "abstract": "Charm production gives rise to a flux of very high energy neutrinos from astrophysical sources with jets driven by central engines, such as gamma ray bursts or supernovae with jets. The neutrino flux from semi-leptonic decays of charmed mesons is subject to much less hadronic and radiative cooling than the conventional flux from pion and kaon decays and therefore has a dominant contribution at higher energies, of relevance to future ultrahigh energy neutrino experiments. ", "introduction": "Large underground or underwater experiments like IceCube~\\cite{Ahrens:2003ix} and KM3NeT~\\cite{Katz:2006wv} are designed with the goal of observing high energy neutrinos produced in astrophysical sources. The highest energy neutrinos, with energies of $10^{9}$~GeV and higher may be observed in radio detection experiments \\cite{radio}, and with an even higher energy threshold of $10^{12}$~GeV with acoustic detection experiments ~\\cite{Vandenbroucke:2004gv}. We consider astrophysical sources driven by a relativistic jet outflow, accelerated by a central engine such as a black hole~\\cite{Woosley:1993wj,grbreview}. Shock accelerated protons in the jet outflow may give rise to a high-energy neutrino flux~\\cite{Waxman:1997ti}. These neutrinos are potentially produced in hadronic interactions: proton--proton interactions produce charged pions and kaons which subsequently decay into muons and neutrinos. Above the threshold for $\\Delta^+$ production, proton interactions with ambient photons also produce charged pions, and at higher energies, kaons. The relative importance of the $pp$ and $p\\gamma$ contributions to the neutrino fluxes depends on the characteristics of the astrophysical environment. Several types of astrophysical sources have been studied in e.g.~\\cite{Waxman:1997ti,RMW,Ando:2005xi,Koers:2007je,Meszaros:2001ms,Razzaque:2003uv,Horiuchi:2007xi,Wang:2008zm,Murase:2008sp,kt1,kt2}. High energy pions and kaons are relatively long-lived and therefore subject to both hadronic and radiative cooling before they decay, which downgrades the neutrino energies. Charm production and decay in astrophysical jets is also a source of neutrinos \\cite{kt1}. In this paper, we show that production of charmed mesons in $pp$ collisions gives a large contribution to the neutrino flux at the highest energies, since high energy charmed hadrons ($D^\\pm, D^0$) have short lifetimes and therefore predominantly decay before they interact. Moreover, since the amount of radiative cooling scales as $m^{-4}$, the larger masses of the charmed hadrons lead to less cooling. The neutrino flux from charm is therefore less suppressed up to higher energies. Even though the production cross section is orders of magnitude smaller than for pions and kaons, neutrinos from charm decays become the dominant contribution at high energies. The energy at which charm begins to dominate depends on the detailed properties of the astrophysical source, and the extent to which charm contributes depends on the maximum energy of the accelerated protons in the jet. An example of an astrophysical source model is the slow-jet supernova (SJS) model, proposed by Razzaque, M\\'esz\\'aros and Waxman (RMW) \\cite{RMW}. It is characterized by a mildly relativistic jet propagating in a collapsing star. This jet does not emerge from the source, and is sometimes referred to as a `choked jet.' This environment has a large optical depth \\cite{Ando:2005xi,Koers:2007je}, so neutrinos may be the only high energy signals to emerge. On the other hand, the jet in a Gamma Ray Burst (GRB) (see e.g.~\\cite{grbreview} for a review) is highly relativistic and the optical depth might be such that photons escape during the burst and do not thermalize. Other GRBs may be choked, so that only neutrinos escape~\\cite{Meszaros:2001ms}. We present here a new analysis of the neutrino flux for two types of source environments: the SJS and a GRB model, both of which have large magnetic fields in the jets. Our treatment of the proton--proton collisions accounts for charmed meson production, as well as pion and kaon production. The shorter lifetimes of charmed mesons allow this channel to potentially dominate the neutrino flux from $pp$ collisions. ", "conclusions": "" }, "0808/0808.2346_arXiv.txt": { "abstract": "The dynamical effects of magnetic fields in models of radiative, Herbig-Haro (HH) jets have been studied in a number of papers. For example, magnetized, radiative jets from variable sources have been studied with axisymmetric and 3D numerical simulations. In this paper, we present an analytic model describing the effect of a toroidal magnetic field on the internal working surfaces that result from a variability in the ejection velocity. We find that for parameters appropriate for HH jets the forces associated with the magnetic field dominate over the gas pressure force within the working surfaces. Depending on the ram pressure radial cross section of the jet, the magnetic field can produce a strong axial pinch, or, alternatively, a broadening of the internal working surfaces. We check the validity of the analytic model with axisymmetric numerical simulations of variable, magnetized jets. ", "introduction": "It is now relatively certain that some Herbig-Haro (HH) jets have knot structures which are the result of a time-variability in the ejection. For example, the observations of some jets with organized structures of knots of different sizes (e.~g., HH 30, 34 and 111, see \\citealt{esq07}, \\citealt{rag02} and \\citealt{mas02}) can be reproduced surprisingly well with variable ejection jet models. In the present paper, we study the effect of the presence of a magnetic field on the evolution of a variable jet. It is still an open question to what extent magnetic fields are important in determining the dynamics of HH jets. The associated problem of radiative, MHD jets has been explored in some detail in the existing literature. \\citet{cer97}, and \\citet{cer99} computed 3D simulations of radiative, MHD jets with different magnetic field configurations (at the injection point). \\citet{fra98} carried out axisymmetric simulations of similar flows. The problem of an MHD, radiative jet ejected with a time-variable velocity was explored with axisymmetric simulations by \\citet{gara00, garb00, sto00, osu00, fra00, dec06} and \\citet{har07}. Variable, MHD jets were also explored with 3D simulations by \\citet{cer01a,cer01b}. The general conclusions that can be obtained from these simulations is that the internal working surfaces produced by the ejection variability are not affected strongly by a poloidal magnetic field. On the other hand, if the magnetic field is toroidal (or, alternatively, has a strong toroidal component), the material within the working surfaces of the jet flow has a stronger concentration towards the jet axis. \\citet{gara00} showed that in a variable ejection velocity jet the ``continuous jet beam'' sections in between the working surfaces have a low toroidal magnetic field, which grows in strength quite dramatically when the material goes through one of the working surface shocks into one of the knots. In the present paper, we present a simple, analytic model from which we obtain the conditions under which the toroidal magnetic field produces an axial compression of the internal working surfaces. This analytic model is presented in \\S 2. In \\S 3, we present axisymmetric numerical simulations in which we compare the working surfaces with and without a toroidal magnetic field, showing the effect described by the analytic model. Finally, in \\S 4 we present our conclusions. ", "conclusions": "It is a known result that internal working surfaces in radiative, MHD jets with a toroidal magnetic field configuration form dense, axial structures, which do not appear in unmagnetized jets. We present a simple, analytic model with which we show that the strong jump conditions (applied to one of the working surface shocks) imply that the magnetic force dominates over the gas pressure force within a radiative working surface and that the gas pressure force is dominant for a non-radiative working surface (provided that one has a shock Mach number of at least $M_w\\sim 10$ and an Alv\\'enic Mach number which does not exceed $M_w^2$). Interestingly, the radial dependence of the toroidal magnetic field within a radiative working surface depends only on the cross section of the pre-shock ram pressure $p_{ram}(r)=\\rho(r)v^2(r)$ impinging on the shocks. From equation (\\ref{ffm}), we can see that if we have a $p_{ram}(r)$ that decreases towards the edge of the jet faster than $1/r^2$, the magnetic force within the working surface will be directed outwards, and will tend to increase the width of the working surface. We have run four simulations (with a top hat cross section for $p_{ram}$, that results in an axially directed magnetic pinch within the working surfaces), therefore in complete consistency with our analytic model. We find that in the non-radiative case the presence of a toroidal magnetic field has very little effect on the structure of the internal working surfaces. We also find that for the radiative case, the presence of a toroidal magnetic field produces a strong axial compression of the material within the internal working surfaces (see Figure 4). The analytic model presented in this paper can then be used to decide what ram pressure and toroidal magnetic field cross section to use in a magnetized, radiative, variable jet simulation in order to produce internal working surfaces that show narrower or broader structures than what is obtained in non-magnetized jet simulations. This might be a valuable tool when trying to model the knots in specific HH jets, and might provide a possible method for constraining the strength and the configuration of magnetic fields within such objects." }, "0808/0808.2988_arXiv.txt": { "abstract": "The shape of the primordial matter power spectrum encodes critical information on cosmological parameters. At large scales, in the linear regime, the observable galaxy power spectrum $\\pobs(k)$ is expected to follow the shape of the linear matter power spectrum $\\plin(k)$, but on smaller scales the effects of nonlinearity and galaxy bias make the ratio $\\pobs(k)/\\plin(k)$ scale-dependent. We develop a method that can extend the dynamic range of the primordial matter power spectrum recovery, taking full advantage of precision measurements on quasi-linear scales, by incorporating additional constraints on the galaxy halo occupation distribution (HOD) from the projected galaxy correlation function $\\wpp$. We devise an analytic model to calculate observable galaxy power spectrum $\\pobs(k)$ in real-space and redshift-space, given $\\plin(k)$ and HOD parameters, and we demonstrate its accuracy at the few percent level with tests against a suite of populated $N$-body simulations. Once HOD parameters are determined by fitting $\\wpp$ measurements for a given cosmological model, galaxy bias is completely specified, and our analytic model predicts both the shape and normalization of $\\pobs(k)$. Applying our method to the main galaxy redshift samples from the Sloan Digital Sky Survey (SDSS), we find that the real-space galaxy power spectrum follows the shape of the nonlinear matter power spectrum at the $1-2\\%$ level up to $k=0.2\\hmpci$ and that current observational uncertainties in HOD parameters leave only few percent uncertainties in our scale-dependent bias predictions up to $k=0.5\\hmpci$. These uncertainties can be marginalized over in deriving cosmological parameter constraints, and they can be reduced by higher precision $\\wpp$ measurements. When we apply our method to the SDSS luminous red galaxy (LRG) samples, we find that the linear bias approximation is accurate to 5\\% at $k\\leq0.08\\hmpci$, but the strong scale-dependence of LRG bias prevents the use of linear theory at $k\\geq0.08\\hmpci$. Our HOD model prediction is in good agreement with the recent SDSS LRG power spectrum measurements at all measured scales ($k\\leq0.2\\hmpci$), naturally explaining the observed shape of $\\pobs(k)$ in the quasi-linear regime. The phenomenological ''$Q$-model'' prescription is a poor description of galaxy bias for the LRG samples, and it can lead to biased cosmological parameter estimates when measurements at $k\\geq0.1\\hmpci$ are included in the analysis. We quantify the potential bias and constraints on cosmological parameters that arise from applying linear theory and $Q$-model fitting, and we demonstrate the utility of HOD modeling of high precision measurements of $\\pobs(k)$ on quasi-linear scales, which will be obtainable from the final SDSS data set. ", "introduction": "\\label{sec:int} In the linear regime, the power spectrum of matter fluctuations encodes information about the physics of early universe (e.g., the potential of the field that drives inflation) and about the matter and energy contents of the cosmos. The power spectrum of galaxies can be biased relative to the power spectrum of matter \\citep{nick,BBKS}, but fairly general theoretical arguments imply that the shape of galaxy power spectrum should approach the shape of the linear matter power spectrum $\\plin(k)$ at sufficiently large scales, i.e., \\beeq \\pR(k)=b_0^2~\\plin(k)+N_0, \\label{eq:bias0} \\eneq where $b_0$ is a constant galaxy bias factor and $\\pR(k)$ denotes the real-space galaxy power spectrum \\citep{peter,fry,dhw95,mann,bob2,vijay,andreas,alexia}. The additive ``shot noise'' term $N_0$ reflects both galaxy discreteness and small scale clustering \\citep{bob2,patMc,smith2006}; in general, it can differ from a simple Poisson sampling correction. In the linear regime, distortions of redshift-space structure by peculiar velocities also alter the amplitude but not the shape of galaxy power spectrum \\citep{nick2}. There have therefore been great efforts to measure the galaxy power spectrum on large scales from angular catalogs and redshift surveys and to use the results to test cosmological models (e.g., \\citealt{yupee,baum,feldman,cbp,lcrs,pscrs}). The enormous size of the Two Degree Field Galaxy Redshift Survey (2dFGRS; \\citealt{colless}) and the Sloan Digital Sky Survey (SDSS; \\citealt{sloan}) relative to earlier samples allows much higher precision measurements. The current state-of-the-art power spectrum measurements are Cole~et~al.'s (2005) analysis of the power spectrum in the 2dFGRS, \\citet{nikhil}, and Blake~et~al.'s (2007) analyses of luminous red galaxies (LRGs) with photometric redshifts in the SDSS, and \\citet{will2} and Tegmark~et~al.'s (2006) measurements from the SDSS redshift survey of main sample galaxies and LRGs. This paper investigates the problem of going from the galaxy power spectrum to the linear matter power spectrum, and hence to cosmological conclusions. The latest observational analyses yield impressive statistical precision on scales near the transition from the linear to the nonlinear regime, e.g., typical 1-$\\sigma$ errors of 5$-$10\\% in $P(k)$ at $k\\simeq0.15\\hmpci$. The critical uncertainty in cosmological interpretation is therefore the accuracy of equation~(\\ref{eq:bias0}) on these scales. The effects of nonlinearity and redshift-space distortions on the $matter$ power spectrum can be computed using numerical simulations or tuned analytic models \\citep[and references therein]{smith}, but details of galaxy formation physics can influence the relation between galaxy and matter power spectra in this regime. \\citet{will2} find that linear theory fits imply different cosmological parameters if applied to measurements with $k\\leq0.06\\hmpci$ or with $k\\leq0.15\\hmpci$, indicating that nonlinear effects have become significant in this regime. Furthermore, \\citet{asp} analyze the SDSS and 2dFGRS galaxy samples and find that the measured shapes of galaxy power spectra differ at a level that cannot be explained by the expected cosmic variance. They show that the likely source of the discrepancy is different scale-dependence of galaxy bias, originating from the different color distributions of galaxies in the SDSS and 2dFGRS samples. \\citet{2dfn}, \\citet{max2} and \\citet{nikhil} approach this problem by fitting a parametrized model of scale-dependent bias, \\beeq P_\\up{gal}(k)=b_0^2\\plin(k){1+Qk^2\\over1+Ak}, \\label{eq:qmodel} \\eneq where we use $P_\\up{gal}(k)$ to represent the galaxy power spectrum, which can be either in real-space or redshift-space. The functional form is devised for convenience to approximate the scale-dependent bias of galaxy samples obtained by populating the Hubble volume simulation \\citep{hubblevol} using a semi-analytic model of galaxy formation \\citep{andrew1}. Here $A=1.4\\hmpc$ or $1.7\\hmpc$ for real-space power spectrum $\\pR(k)$ or angle-averaged redshift-space power spectrum $P_0(k)$ measurements, respectively, and $Q$ is treated as a free parameter that is marginalized over in deriving cosmological parameter constraints. This approach is adequate {\\it if} equation~(\\ref{eq:qmodel}) is a sufficiently accurate description of scale-dependent bias for some value of $Q$, but it could yield biased parameter estimates or incorrect error bars if the actual scale-dependence is different. It also gives up on extracting cosmological information from scales where bias might be mildly scale-dependent. For example, \\citet[hereafter, T06]{max2} find that cosmological parameters remain unaffected by changes in power spectrum measurements at $k\\geq0.1\\hmpci$ once they marginalize over the value of $Q$. This implies that the statistical constraining power on cosmological parameters is lost at $k\\geq0.1\\hmpci$ by the marginalization process. In this paper, we present an alternative approach to recovering the shape of the linear matter power spectrum, both more aggressive and more robust than ``marginalizing over $Q$.'' Our approach is based on the halo occupation distribution (HOD) framework, which describes the nonlinear relation between galaxies and matter by specifying the probability $P(N|M)$ that a halo of mass $M$ hosts $N$ number of galaxies of a given type, together with specification of the relative spatial and velocity distributions of galaxies within halos.\\footnote{Throughout this paper, the term ``halo'' refers to a dark matter structure of overdensity $\\rho/\\bar \\rho_m\\simeq200$, in approximate dynamical equilibrium.} The HOD formalism has emerged as a powerful method of modeling galaxy bias \\citep{jing2,uros,chung,john2,roman,andreas} because the dynamics of dark matter halos can be accurately calculated using analytic approximations or $N$-body simulations, and the effects of galaxy formation physics can be parametrized in terms of an HOD and inferred by fitting observational data. Our strategy for extending recovery of the primordial matter power spectrum is to use complementary information from the measurements of the projected correlation function $\\wpp$ as a constraint to obtain HOD parameters given a cosmological model. We then predict the galaxy power spectrum $P_\\up{gal}(k)$ and study the scale-dependent bias \\beeq b^2(k)\\equiv P_\\up{gal}(k)/\\plin(k). \\label{eq:scale} \\eneq For each cosmological model, fitting $\\wpp$ measurements determines HOD parameters and we can then compute a {\\it unique} prediction of $P_\\up{gal}(k)$, both shape and normalization (which is essentially pinned to the amplitude of $\\wpp$). Uncertainties in HOD parameters introduce uncertainty in $P_\\up{gal}(k)$ and $b^2(k)$, but these uncertainties can be accurately computed and marginalized over. Therefore, we can extend the wavenumber range over which $P_\\up{gal}(k)$ measurements can be used for cosmological parameter constraints, taking full advantage of precision measurements on quasi-linear scales. In practice, we are just using the measured $P_\\up{gal}(k)$ and $\\wpp$ to simultaneously constrain HOD parameters and the cosmological parameters and marginalizing over the former. Relative to the $Q$-model approach, our method adopts a more physically motivated computation of $P_\\up{gal}(k)$ and $b^2(k)$, requiring only the validity of the adopted HOD parametrization, and it brings in the additional information present in $\\wpp$ rather than using only the $P(k)$ shape itself to constrain the scale-dependence of bias. In principle, the power spectrum $P(k)$ and correlation function $\\xi(r)$ contain the same information. However, they are in practice measured via different estimators and on different scales, where their signal-to-noise ratios are highest and systematic errors are relatively well understood. The information in $P(k)$ and $\\xi(r)$ measurements on these non-overlapping scales is therefore not identical, but complementary. Furthermore, the projected correlation function $\\wpp$ is measured to ease the difficulty in interpreting nonlinear redshift-space distortion of correlation function measurements on small scales. Therefore, the addition of $\\wpp$ measurements at $r_p\\leq30\\hmpc$ brings new information that is not present in $P(k)$ measurements at $k\\leq1\\hmpci$. A different approach to this problem is to develop an analytic model for predicting the scale-dependence of galaxy bias by using higher-order perturbation theory (e.g., \\citealt{patMc,smith2006}). This approach is elegant and transparent in nature, since it is based on linear theory and its extension to higher-order, while our approach is less {\\it ab initio} in the sense of incorporating elements calibrated by numerical $N$-body simulations in our analytic model. However, the critical uncertainty for this approach based on higher-order perturbation theory is its applicability on quasi-linear scales ($\\gtrsim0.1\\hmpci$), where first-order linear theory is known to be inaccurate, but the measurement precision is highest in practice. In contrast, our approach is fully nonlinear, and phenomenological in nature, so it can be applied down to small scales, limited only by the point at which uncertainties in the HOD parameters introduce systematic uncertainty in the $P(k)$ recovery. Analyses of galaxy redshift surveys typically estimate the angle-averaged power spectrum $P_0(k)$, i.e., the monopole of the redshift-space power spectrum (e.g., \\citealt{2dfn,will2}). Redshift-space distortions do not alter the shape in linear theory, but they do change the shape in the trans-linear regime (e.g., \\citealt{cole1}), and finger-of-god (FoG) effects have impact out to large scales (e.g., \\citealt{roman}). \\citet{nikhil} and \\citet{blake} deproject the angular clustering measurements of the SDSS LRG sample using photo-$z$ catalogs to estimate the real-space power spectrum, independent of redshift-space distortions. \\citet{max1,max2} use a linear combination of the redshift-space monopole, quadrupole, and hexadecapole that recovers the real-space power spectrum in the linear regime. We will denote this ``pseudo real-space'' power spectrum $\\pzr(k)$. The redshift-space power spectrum estimators can be applied directly to galaxy redshift data or applied after compressing FoG effects. We will investigate $P_\\up{gal}(k)$ and $b^2(k)$ for all of these cases. To this end, we develop an analytic model in \\S~\\ref{sec:method} for calculating real-space and redshift-space galaxy power spectra given $\\plin(k)$ and a galaxy HOD, drawing on the \\citet{jeremy2} model for redshift-space distortion, which improves on previous work (e.g., \\citealt{uros3,martin,kang,coor}). \\citet{jeremy2} tests the model for computing the redshift-space correlation function against a series of populated $N$-body simulations. Here we extend the model and present additional tests of its applicability to modeling redshift-space power spectra in \\S~\\ref{sec:gal}. In this paper, we use HOD parameters for volume-limited galaxy samples that have well defined classes of galaxies, focusing on SDSS main galaxy samples with absolute-magnitude limits $M_r\\leq-20$ and $M_r\\leq-21$ \\citep{idit3} in \\S~\\ref{sec:lin}, and SDSS LRG samples with absolute-magnitude limits $-23.2\\leq M_g\\leq-21.2$ and $-23.2\\leq M_g\\leq-21.8$ \\citep{daniel,idit4,zheng5} in \\S~\\ref{sec:lrg} for application of our method.\\footnote{For brevity, we quote the absolute magnitude thresholds $M_r-5\\log h$ and $M_g-5\\log h$ for $h\\equiv1$.} More complete modeling of the conditional luminosity function \\citep{yang} might allow use of flux-limited galaxy catalogs, though it requires more free parameters to provide complete descriptions of the galaxy samples. Here we only consider volume-limited galaxy samples, whose results can be combined to improve statistical precision. We summarize our main results in \\S~\\ref{sec:sum}. ", "conclusions": "\\label{sec:sum} We have developed an analytic model to predict observable galaxy power spectra $\\pobs(k)$ for specified cosmological and galaxy HOD parameters, and we have verified its accuracy using $N$-body simulations. As potentially observable power spectra $\\pobs(k)$, we have considered the real-space $\\pR(k)$, the redshift-space monopole $P_0(k)$, and the pseudo real-space $\\pzr(k)$, with varying levels of Finger-of-God (FoG) compression for the latter two. Once HOD parameters are determined by fitting the number density $\\bar n_g$ and projected correlation function $\\wpp$ of the observed SDSS galaxy samples, given a specified cosmological model, our analytic model can be used to predict $\\pobs(k)$ measurements. The large-scale normalization of our predictions is also fixed in the process of fitting $\\wpp$, providing a unique prediction for each combination of cosmological and HOD parameters. In practice, one can simultaneously fit cosmological and HOD parameters using $\\pobs(k)$ and $\\wpp$ as constraints, then marginalize over the HOD in deriving cosmological parameters. By implementing a complete physical model of nonlinear galaxy bias and drawing on the additional information in $\\wpp$, our method allows one to take full advantage of precision measurements of $\\pobs(k)$ on quasi-linear scales ($k=0.1-0.4\\hmpci$), where linear theory or the phenomenological $Q$-model may be insufficiently accurate. Our main findings are as follows: 1. Our analytic model for calculating $\\wpp$ follows the method described in \\citet{jeremy}, with the improved treatment of the scale-dependent halo bias and ellipsoidal halo exclusion corrections. Drawing on the \\citet{jeremy2} model for redshift-space distortion, the analytic model is extended to incorporate calculating real-space and redshift-space power spectra. We have tested its predictions for $\\wpp$ and $\\pobs(k)$ against populated $N$-body simulations spanning cosmological parameter range $\\OM=0.1-0.63$ and $\\rms=0.6-0.95$, with HOD parameters matched to represent two SDSS galaxy samples with absolute-flux limits $M_r\\leq-20$ and $M_r\\leq-21$ \\citep{idit3}. The analytic model reproduces the numerical results of $\\wpp$ to 5\\% or better, and the predictions of $\\pobs(k)$ are consistent with the numerical results to 2\\% at $k=0.1-1\\hmpci$ and to 10\\% at $k=0.025-0.1\\hmpci$, though the finite box size of the simulations makes it difficult to assess the statistical significance of differences on large scales. 2. For the $M_r\\leq-20$ galaxy sample, the pseudo real-space power spectrum $\\pzr(k)$ recovers the true $\\pR(k)$ to 2\\% at $k\\leq0.2\\hmpci$, while the deviation between $\\pR(k)$ and the scaled monopole $P_0(k)$ is already 10\\% at $k=0.1\\hmpci$. However, the deviation of $\\pzr(k)$ from $\\pR(k)$ becomes substantial at $k\\geq0.3\\hmpci$. This deviation can be partly remedied by FoG compression, which suppresses nonlinear behavior of the redshift-space multipoles caused by the random motions of satellite galaxies within halos. With FoG compression threshold $\\sigma_\\up{h}=750~\\kms$, $\\pzrF{750}(k)$ can recover $\\pR(k)$ to 5\\% at $k\\leq0.45\\hmpci$, and at higher $k$ for $\\pzrF{400}(k)$. FoG compression also reduces nonlinearity of the monopole power spectrum, but $P_0^{400}(k)$ can only achieve 10\\% accuracy at $k\\leq0.3\\hmpci$. We conclude that the pseudo real-space method of \\citet{max0} is an effective tool for recovering the nonlinear real-space galaxy power spectrum from redshift-space measurements, especially if it is combined with accurate FoG compression. 3. The nonlinear $matter$ power spectrum describes the nonlinear real-space galaxy power spectra to 1\\% at $k\\leq0.2\\hmpci$ for the $M_r\\leq-20$ and $M_r\\leq-21$ galaxy samples, up to an overall bias factor $b^2_0$. The shape of the scale-dependent bias function $b^2(k)/b_0^2$ for $\\pzr(k)$ is qualitatively similar to $\\pR(k)$ at $k\\leq0.3\\hmpci$, but the shape for $P_0(k)$ is completely different over the entire range we consider here. FoG compression makes little difference to $b^2(k)/b^2_0$ for $\\pzr(k)$, but a large difference for $P_0(k)$. For these SDSS main galaxy samples, the $Q$-model prescription traces our calculation of $\\pR(k)$ relatively well at $k\\geq0.1\\hmpci$, but its shape on large scales differs, so it might induce some overall bias in cosmological parameters when fitted to $\\pobs(k)$ measurements that have large uncertainties at $k\\leq0.05\\hmpci$. Similar trends but with larger discrepancy are found in comparison to our $\\pzr(k)$ and $P_0(k)$ calculations. 4. Uncertainties in computing $\\pobs(k)$ in our method arise from observational uncertainties in the HOD parameters and from uncertainty in the adopted parametrization itself. We have examined these uncertainties by adopting a flexible HOD parametrization with freedom to explore a wider range of plausible halo occupation functions. For the $M_r\\leq-20$ sample with the \\citet{idit3} uncertainties in $\\wpp$, the uncertainty in the predicted $\\pobs(k)$ is 2\\% at $k=0.2\\hmpci$, becomes progressively smaller at lower $k$, and climbs up to 4\\% at $k=0.5\\hmpci$. The uncertainty is a factor of two smaller for the $M_r\\leq-21$ sample, roughly the ratio of the fractional $\\wpp$ measurement errors of the two samples. We have not investigated the uncertainties associated with possible environmental variations of the HOD \\citep{croton}. Based on work to date, we expect that such variations might lead to few percent uncertainties in the overall normalization predicted for $\\pobs(k)$ after fitting $\\wpp$, but that the impact on scale-dependence of $b^2(k)/b^2_0$ would be smaller. 5. Moving to the LRG regime, we have tested our analytic model predictions against the $z=0$ output of a large volume, 1024$^3$-particle $N$-body simulation \\citep{warren1}, populated based on Zheng et~al.'s (2008) HOD fits to $\\wpp$ for two volume limited SDSS LRG samples \\citep{idit4}. The analytic model predicts $\\xi_0(r)$ and $\\xiR(r)$ to 5\\% or better over the range $0.1\\hmpc\\leq r\\leq30\\hmpc$, and the predictions for $P_0(k)$ and $\\pR(k)$ have similar accuracy over the range $0.01\\hmpci\\leq k\\leq0.3\\hmpci$. 6. For the LRG samples, the linear (scale-independent) bias approximation remains accurate at the 5\\% level to $k=0.08\\hmpci$ for the $-23.2\\leq M_g\\leq-21.2$ sample and to $k=0.1\\hmpci$ for the $-23.2\\leq M_g\\leq-21.8$ sample. There is little variation among $\\pzr(k)$, $P_0(k)$, and $\\pR(k)$, because LRGs are mainly central galaxies in massive halos, so random motions of satellite galaxies have little impact. Similarly, FoG compression has only a small impact on $b^2(k)/b_0^2$ for these samples. Both samples show strong scale-dependence of bias at $k\\geq0.1\\hmpci$, much more than for main sample galaxies.\\footnote{This result, obtained by fitting HODs and computing $\\pobs(k)$ with our analytic model, confirms the result of T06 inferred by fitting $\\pobs(k)$ with $Q$-models. However, the actual scale-dependence we find for LRGs is somewhat weaker than that inferred by T06.} If we fit $b^2(k)/b_0^2$ from our HOD models with the best $Q$-model over the range $k=0.01-0.2\\hmpci$, the largest deviation is 7\\%. 7. We have presented a preliminary comparison of our analytic model predictions to the T06 measurements of the LRG $\\pobs(k)$, with no FoG compression. The difference between our volume-limited samples and the T06 flux-limited sample precludes a full quantitative assessment, but the qualitative agreement is remarkably good over the full range of the measurements, $k=0.01-0.2\\hmpci$ (Fig.~\\ref{fig:lrg-pow}). Fits with different cosmological parameters differ on large scales and, to a smaller degree, at $k\\geq0.2\\hmpci$, indicating that measurements to smaller scales would provide additional discriminatory power. 8. Looking to the future, we have generated synthetic $\\pobs(k)$ data from our analytic model with error bars half those of T06, then fit them to successively higher $k_\\up{max}$ with linear theory, the $Q$-model, and the HOD model. Cosmological parameters from linear theory fits are badly biased for $k_\\up{max}\\geq0.1\\hmpci$, while for $k_\\up{max}=0.09\\hmpci$ they are biased at less than 1-$\\sigma$. Parameters from the $Q$-model are minimally biased for $k_\\up{max}=0.09\\hmpci$, biased by 1.2-$\\sigma$ for $k_\\up{max}=0.2\\hmpci$, and biased by many-$\\sigma$ for $k_\\up{max}=0.4\\hmpci$. Since the synthetic data are generated from the HOD model, the HOD parameter estimates are unbiased, and the error bars in cosmological parameters shrink steadily as $k_\\up{max}$ is increased from $0.1\\hmpci$ to $0.2\\hmpci$ to $0.4\\hmpci$. Results~7 and~8 are especially encouraging. Using only the HOD model and the information in $\\wpp$, our method predicts exactly the scale-dependent bias for LRGs that is required to transform the linear power spectrum from WMAP3 into the SDSS galaxy power spectrum measured by T06. This is in contrast to $Q$-model fitting, where a phenomenological parameter (motivated by simulation results but with no clear physical interpretation) is introduced specifically to account for the difference between the linear theory $P(k)$ and the observed power spectrum. Despite the clear evidence that scale-dependent bias affects the LRG power spectrum beyond $k=0.1\\hmpci$, result~8 shows that one can gain substantial additional leverage on cosmological parameters with HOD modeling of power spectrum measurements up to $k=0.2-0.4\\hmpci$, and possibly beyond. Realizing this opportunity will require several investigations beyond those presented here. First, we will need more large volume simulations to test and, if necessary, refine our analytic model to the level of accuracy demanded by the final SDSS data set. Second, we must explore more thoroughly the uncertainties associated with the HOD fitting, including alternative parametrizations, the impact of velocity bias on redshift-space predictions, and the possible impact of environmental variations of the HOD. Given the growth of current and future galaxy surveys in depth and redshift, these investigations will be needed to go beyond linear theory. Precise measurement of the primordial matter power spectrum will play a crucial role in constraining cosmological parameters and testing dark energy theories." }, "0808/0808.2493_arXiv.txt": { "abstract": "Studies of strong gravitational lensing in current and upcoming wide and deep photometric surveys, and of stellar kinematics from (integral-field) spectroscopy at increasing redshifts, promise to provide valuable constraints on galaxy density profiles and shapes. However, both methods are affected by various selection and modelling biases, whch we aim to investigate in a consistent way. In this first paper in a series we develop a flexible but efficient pipeline to simulate lensing by realistic galaxy models. These galaxy models have separate stellar and dark matter components, each with a range of density profiles and shapes representative of early-type, central galaxies without significant contributions from other nearby galaxies. We use Fourier methods to calculate the lensing properties of galaxies with arbitrary surface density distributions, and Monte Carlo methods to compute lensing statistics such as point-source lensing cross-sections. Incorporating a variety of magnification bias modes lets us examine different survey limitations in image resolution and flux. We rigorously test the numerical methods for systematic errors and sensitivity to basic assumptions. We also determine the minimum number of viewing angles that must be sampled in order to recover accurate orientation-averaged lensing quantities. We find that for a range of non-isothermal stellar and dark matter density profiles typical of elliptical galaxies, the combined density profile and corresponding lensing properties are surprisingly close to isothermal around the Einstein radius. The converse implication is that constraints from strong lensing and/or stellar kinematics, which are indeed consistent with isothermal models near the Einstein radius, cannot trivially be extrapolated to smaller and larger radii. ", "introduction": "\\label{S:motivation} \\subsection{Learning from galaxy density profiles and shapes} \\label{SS:intro1} Observational constraints on galaxy density profiles and shapes can be used to study a great variety of problems, from basic cosmology to the connections between dark matter and baryons. For example, the spherically-averaged form of the dark matter halo density profile, and the distribution of halo shapes, both depend on cosmological parameters to the extent that those parameters affect the process of structure formation \\citep[e.g.,][]{2001MNRAS.321..559B, 2006MNRAS.367.1781A}, and on the physics of galaxy formation to the extent that baryons modify the dark matter distribution \\citep[e.g.,][]{1986ApJ...301...27B, 2004ApJ...616...16G, 2004ApJ...611L..73K, 2007ApJ...658..710N, 2008ApJ...672...19R}. While there is still some disagreement in the literature about the theoretical predictions, even among those papers cited above, one key result is that there is considerable scatter in both the density profiles and shapes of dark matter halos, which presumably reflects different formation histories \\citep[e.g.,][]{2002ApJ...568...52W}. Several observational tools have been used to constrain the density profiles and shapes of galaxies, at different length and mass scales, and redshifts. The main tools we consider are gravitational lensing and stellar kinematics. X-ray data are also useful for understanding the density profiles and shapes of massive galaxies and galaxy clusters, but are beyond the scope of our investigation. Gravitational lensing is the deflection of light from distant sources by the gravitational fields of intervening lens galaxies. Strong lensing occurs in the central regions of galaxies (and clusters) where the light bending is extreme enough to produce multiple images of a background source. It can be used to study the mass distributions of individual galaxies on projected scales of typically several kpc \\citep[for a review, see ][]{Saas-Fee}. Weak lensing is a complementary phenomenon in which the deflection of light slightly distorts the shapes of background sources without creating multiple images \\citep[for a review, see][]{2001PhR...340..291B}. Weak lensing probes galaxy mass distributions on scales from tens of kpc out to about 10 Mpc, but only yields constraints on ensemble averages, since currently weak lensing by galaxies can only be detected by stacking many lens galaxies. While galaxy clusters can by studied on an individual basis using weak lensing, our focus is on galaxies. High-quality (two-dimensional) data on the kinematics of stars as well as gas in the inner parts (a few to ten kpc) of galaxies are now readily available, thanks in particular to strong progress in integral-field spectroscopy in the last decade. Kinematics in the outer parts of late-type galaxies can often be observed from the presence of neutral hydrogen \\citep[e.g.,][]{1981AJ.....86.1825B, 1985ApJ...295..305V, 1996MNRAS.281...27P, 2007MNRAS.376.1513N}. In the outer parts of early-type galaxies, however, cold gas is scarce \\citep[but see e.g.][]{1994ApJ...436..642F, 1997AJ....113..937M, 2008MNRAS.383.1343W}, so we are left with discrete kinematic tracers such as planetary nebulae and globular cluster \\citep[e.g.,][]{2003ApJ...591..850C, 2007ApJ...664..257D} out to tens of kpc. Current and upcoming wide and deep photometric surveys will reveal a vast number of strong lensing events \\citep[e.g.,][]{ 2004AAS...20510827F, 2004NewAR..48.1085K, 2004ApJ...601..104K, 2005NewAR..49..387M}, and will also allow for extensive, complementary weak lensing analysis \\citep[e.g.][]{2002SPIE.4836...10T, 2004SPIE.5489...11K}. At the same time, there is a rapid increase in the availability of two-dimensional kinematics of galaxies nearby \\citep[e.g.,][]{2004MNRAS.352..721E, 2006MNRAS.373..906M}, and even at high(er) redshift \\citep[e.g.,][]{2007ApJ...668..738V, 2007ApJ...671..303B}. These data sets, individually as well as combined, may provide strong constraints on galaxy density profiles and shapes, but the analyses involved are subject to selection and modelling biases. \\subsection{Selection and modelling biases} \\label{SS:intro2} If we want to use strong lensing and kinematics to make robust tests of cosmological predictions, we need to answer two questions: Can strong lensing and kinematic analyses yield accurate constraints on galaxy density profiles? Are the galaxies in which we can make the measurements (especially in the case of strong lensing) representative of all galaxies? We refer to these two concerns as \\emph{modelling biases} and \\emph{selection biases}, respectively. The question of selection bias is particularly important given the diversity in the galaxy population. It is well known that strong lensing favours early-type over late-type galaxies, and massive galaxies over dwarfs \\citep[e.g.,][]{1984ApJ...284....1T, 1991MNRAS.253...99F}. But even within the population of massive early-type galaxies, to what extent does strong lensing favour galaxies whose dark matter halos have inner slopes that are steeper than average, or concentrations that are higher than average? Also, to what extent does strong lensing favour galaxies with particular shapes and/or orientations with respect to the line-of-sight? To phrase these questions formally, consider some parameter $x$ describing the galaxy density profile or shape (e.g., the inner slope of the dark matter density profile). We need to understand how the distribution $p_\\mathrm{SL}(x)$ among strong lens galaxies compares to the underlying distribution $p(x)$ for all galaxies. Any analysis that includes strong lensing with some other technique used to study the same systems will suffer from strong lensing-related selection biases, since we never get to choose which systems will be a strong lens. Selection biases are critical when we want to interpret constraints on lens galaxy density profiles and/or shapes from strong lensing, possibly in combination with kinematics \\citep[e.g.][]{2005ApJ...623..666R, 2006ApJ...649..599K} in comparison with predictions from cosmological models. On the other hand, modelling biases may occur because of (often unavoidable) assumptions made when analysing gravitational lensing and/or kinematic data because of the finite number of constraints available from the data. Since galaxies are in general non-spherical, we can only correctly interpret the observations if we know the viewing direction, and even then the deprojection might not be unique \\citep[e.g.][]{1987IAUS..127..397R}. Moreover, strong lensing is subject to the so-called mass-sheet degeneracy: part of the deflection and magnification of the light from the background source can be due to mass along the line-of-sight that is not associated with the lens galaxy itself\\footnote{Part of this mass-sheet degeneracy may be overcome in future surveys through the measurements of time delays for an adopted Hubble constant.}. In addition, the constraints from strong lensing on the galaxy density are typically limited in radius to around the Einstein radius, and even then, without secure, unaffected measurements of the flux of the lensing images, the density profile is difficult to recover. Kinematics can be obtained over a larger radial extent, but in particular stellar kinematics suffer from the so-called mass-anisotropy degeneracy: a change in the measured line-of-sight velocity dispersion can be due to a change in total mass, but may also be the result of velocity anisotropy. In a series of papers, we are developing a pipeline for using realistic galaxy models to simulate strong lensing and kinematic data, and assess how both selection and modelling biases affect typical strong lensing and kinematic analyses. In this first paper, we present the simulation pipeline for point-source lensing. We focus on strong lensing of quasars by early-type, central galaxies at different mass scales, using two-component mass profiles (dark matter plus stellar component) that are consistent with existing photometry and stacked weak lensing data from SDSS \\citep{2003MNRAS.341...33K,2006MNRAS.368..715M}, $N$-body, and hydrodynamic simulations. In \\citeauthor{2008paperII} (\\citeyear{2008paperII}, hereafter Paper~II), we use this pipeline to study selection biases in strong lensing surveys. In future work we will address modelling biases in strong lensing and kinematics (both when studied separately and when combined). Our overall goal is to determine how strong lensing, kinematics, and possibly other probes (weak lensing, and X-ray data for cluster mass scales) can be used to obtain robust, unbiased constraints on galaxy density profiles and shapes. To make our presentation coherent, and to clarify our notation and terminology, we first review the analytic treatment of ellipsoidal mass distributions with different profiles and shapes (Section~\\ref{S:profiles-shapes}), and the basic theory of strong lensing (Section~\\ref{S:stronglensing}). We then describe our simulation pipeline for point-source lensing in depth. In Section~\\ref{S:galmodels} we present our choices for the masses, profiles, and shapes of the galaxy models. In Section~\\ref{S:lensingcalc} we discuss the numerical methods we use for lensing calculations, including numerous tests. A summary of the simulation pipeline, and some implications for strong lensing analyses are given in Section~\\ref{S:conclusions}. ", "conclusions": "\\label{S:conclusions} We have presented a flexible simulation pipeline for coherent investigations of selection and modelling biases in strong lensing surveys. We have focused on point-source lensing by two-component galaxy models meant to emulate realistic early-type, central galaxies at two different mass scales: a lower, $\\sim 2L_*$ galaxy mass scale and a higher, $\\sim 7 L_*$ group mass scale (with $L_*$ in the $r$-band). Below is a list of our main conclusions regarding the construction of the simulation pipeline for lensing by realistic galaxy models: \\begin{itemize} \\item We include both cusped and deprojected S\\'ersic density profiles, with observationally-motivated choices of masses and scale lengths, and a range in density profile parameters, separately for the stellar and dark matter component. We use seven different models for the galaxy shapes with the stellar component rounder than the dark matter component, but with the axes intrinsically aligned. [Sections~\\ref{SS:massscales}--\\ref{SS:sersicsim}] \\item When we change the density profile parameters away from the adopted fiducial values, we preserve the total and virial mass of the stellar and dark matter components (respectively) by changing the density amplitude. We keep the scale radius fixed to comply with observed size-luminosity relations for early-type galaxies and concentration-mass relations for dark matter halos. Similarly, for the non-spherical shapes we change the scale length to preserve the mass but keep the scale radius (and hence the concentration) the same as for the fiducial spherical case. [Sections~\\ref{SSS:normalization}, \\ref{SSS:scalelengthnonspherical}, \\ref{SSS:sersicnormalization}] \\item We choose fiducial lens and source redshifts of $0.3$ and $2.0$, resulting in Einstein radii $\\Rein$ at about one effective radius $R_e$. [Section~\\ref{SS:redshifts}] \\item We require multiple ways of handling magnification bias to sample different parts of the observed quasar luminosity function and to allow for application to different survey limitations in image resolution and flux. [Section~\\ref{SS:magbias}] \\item When the surface mass density is known analytically (as for our deprojected S\\'ersic models), we compute the corresponding lensing deflection and magnification analytically (for circular symmetry) or with numerical integrals (for elliptical symmetry). When the surface mass density cannot be computed analytically, as for the cusped models, we construct a surface density map and use Fourier methods to compute the lensing deflection and magnification. We validate and use Monte Carlo methods to calculate lensing statistics, including unbiased and biased cross-sections and image separations. [Sections~\\ref{SS:kapmap}--\\ref{SS:validation}] \\item For our map-based lensing calculations, tests for convergence at the 5 per cent level suggest that the required map resolution is $\\sim 25$ pixel lengths per $\\Rein$, corresponding to $0.025\\arcsec$ ($\\simeq 0.11$\\,kpc) and $0.075\\arcsec$ ($\\simeq 0.33$\\,kpc) for the lower and higher mass scale, respectively. The resolution needs to be fine enough to simultaneously resolve the steepness of the inner slope of the density and recover the smoothness of the shape of the inner critical curve. Similarly, convergence tests suggest that the required box size is $\\sim 16\\,\\Rein$, corresponding to $10\\arcsec$ ($\\simeq 43.6$\\,kpc) and $30\\arcsec$ ($\\simeq 131$\\,kpc) for the lower and higher mass scale, respectively. [Sections~\\ref{SS:resolution} and~\\ref{SS:boxsize}] \\item Lensing cross-sections for fairly general triaxial models vary in a smooth and monotonic way with viewing direction, which can be efficiently sampled through projected axis ratios (rather than viewing angles) of the surface density. Surprisingly few samplings are necessary: 10, 8, and 24 viewing directions can effectively determine the orientation-averaged cross-sections for oblate, prolate, and triaxial models (respectively) with 5 per cent precision. [Section~\\ref{SS:projangle}] \\end{itemize} We have implemented the construction of the realistic galaxy models and efficient computation of the corresponding surface mass density maps in \\idl\\ (see also Section~\\ref{SS:compsurfmaps}). All subsequent strong lensing calculations are done with an updated and extended version of \\gravlens. We use scripts to coordinate and analyse the extensive data flow to and from these two parts, and to make it into a coherent simulation pipeline. Our investigations have revealed a number of useful, general points related to the interpretation of observational strong lensing results: \\begin{itemize} \\item In the vicinity of the Einstein radius, the intrinsic and projected logarithmic slopes of the total density are close to the values $\\gamma=2$ and $\\gamma'=1$ of an isothermal profile, for all of our galaxy models (see Figs.~\\ref{F:3dlogslope} and~\\ref{F:2dlogslope}, and the text of Sec.~\\ref{SSS:intrinsicprofiles}). Consequently, a measurement of the total density profile near the Einstein radius cannot (alone) be used to determine the dark matter inner slope for a two-component model. This is a non-trivial conclusion given the variety of intrinsic non-isothermal density profiles used for both the stellar and dark matter component, including profiles with a true central cusp and those that do not asymptote to a particular inner slope at any scale. \\item We also find that the lensing deflection curves are nearly flat around (and even beyond) $\\Rein$, similar to a flat deflection curve for an isothermal model (see Fig.~\\ref{F:alpha}), particularly for the lower-mass galaxy scale. This again implies that constraints from strong lensing cannot be extrapolated to radii much smaller or larger than the Einstein radius. We emphasize that the parameters for the galaxy models were \\emph{not} a priori chosen to mimic this effect, but that it truly seems to be an ``isothermal conspiracy'' \\citep[e.g.,][]{2003ApJ...595...29R}. \\item Because of our assumption of alignment between the intrinsic axes of the stellar and dark matter components, only the two shape models with a triaxial dark matter component and a rounder (triaxial or oblate) stellar component can have a significant misalignment between their projected axes. Even then, significant misalignment occurs only for a limited number of viewing directions, several of which lead to a relatively round projected stellar component such that the misalignment is not very important in practice (the position angle of a round stellar component is on the sky, after all, difficult to establish). The small misalignments inferred for relatively isolated lens systems, as well as detailed dynamical modelling of early-type galaxies, support near intrinsic alignment between stars and dark matter (see Section~\\ref{SSS:misalignment}). \\end{itemize} In Paper~II, we use the flexibility of the pipeline to study selection biases related to the galaxy mass, shape, orientation, and various parameters of the dark matter and stellar profiles. In subsequent work, we will investigate modelling biases by analysing the mock lens systems produced by the pipeline using the lens modelling tools within the \\gravlens\\ package. We have therefore created a lensing simulation pipeline that will be crucial for deriving inferences about galaxy density profiles and shapes, $H_0$, and other parameters from the thousands of lenses that will be discovered in the coming years in large photometric surveys such as Pan-STARRS, LSST, SNAP, and SKA. Furthermore, following a similar approach as in \\citeauthor{2008MNRAS.385..614V} (2008a), we plan to extend the pipeline to produce projected kinematics of our galaxy models that mimic the two-dimensional kinematic observations of early-type galaxies. We will then investigate how well we can expect to recover the intrinsic density profile and shape of early-type galaxies, based solely on dynamical models fitted to these simulated kinematics, or on a combination of kinematics and strong lensing (using techniques similar to \\citeauthor{2008arXiv0807.4175V} 2008b). This work will be valuable for understanding the modelling biases in analyses of the kinematic data that are becoming available for hundreds of galaxies at increasing redshifts, and for understanding both modelling and selection biases in joint lensing+kinematics studies." }, "0808/0808.1258_arXiv.txt": { "abstract": "{Modelling the emission properties of compact high energy sources such as X-ray binaries, AGN or $\\gamma$-ray bursts represents a complex problem. Contributions of numerous processes participate non linearly to produce the observed spectra: particle-particle, particle-photon and particle-wave interactions. Numerical simulations have been widely used to address the key properties of the high energy plasmas present in these sources.} {We present a code designed to investigate these questions. It includes most of the relevant processes required to simulate the emission of high energy sources. } {This code solves the time-dependent kinetic equations for homogeneous, isotropic distributions of photons, electrons, and positrons. We do not assume that the distribution has any particular shape. We consider the effects of synchrotron self-absorbed radiation, Compton scattering, pair production/annihilation, e-e and e-p Coulomb collisions, e-p bremsstrahlung radiation and some prescriptions for additional particle heating and acceleration.} {We illustrate the code's computational capabilities by presenting comparisons with earlier works and some examples. Previous results are reproduced qualitatively but some differences are often found in the details of the particle distribution. As a first application of the code, we investigate acceleration by second order Fermi-like processes and find that the energy threshold for acceleration has a crucial influence on the particle distribution and the emitted spectrum.}{} ", "introduction": "High energy sources, such as X-ray binaries, active galactic nuclei (AGN hereafter), or $\\gamma$-ray bursts, exhibit spectra detectable to very high energy. This radiation must originate in a plasma for which a significant fraction of the particles have relativistic energies. Understanding the properties of these hot plasmas remains a challenge in the modelling of X- and $\\gamma$-ray sources. Among the many processes at work, there are particle-particle interactions, such as Coulomb collisions, particle-photon interactions such as Compton scattering, synchrotron radiation, bremsstrahlung emission, or pair production/annihilation, and particle-wave interactions that lead to particle acceleration. However, the way they add or compete is highly non-linear and the cross sections involved are complex. Investigating a large parameter space is required, and in spite of important breakthroughs, these plasmas are still poorly understood. Analytical studies provided interesting qualitative results with approximations, but a more general approach based on numerical simulations is required to explain the details and complexity of contemporary observations. The first detailed investigations were analytical attempts to model the Compton scattering in thermal plasmas of fixed temperature \\citep[][]{BZS71,Sunyaev80,Guilbert81,Zdziarski85,Guilbert86}. In parallel, some of these results were confirmed by Monte-Carlo simulations \\citep[e.g.][]{PSS83,Gorecki84}. The additional role of pair production and annihilation in thermal plasmas, whose temperatures were determined self-consistently, was then studied both analytically \\citep{Svensson82b,Svensson83,Guilbert85,Kusunose87} and numerically \\citep{Zdziarski84,Zdziarski85}. These works constituted significant advances because they explicitly accounted for the back reaction of the radiation field on the plasma temperature. However, they were limited to thermal distributions of particles, whereas significant evidence of strongly non-thermal populations was found in many sources. For instance, spectra of blazars or radio loud AGN were shown to be shaped at least by the synchrotron self-compton emission of purely non-thermal electrons \\citep[e.g.][]{Ghisellini98}. At these high energies, accelerated particles cool on very short timescales before they can be thermalized for instance by two body collisions. The balance between this cooling and acceleration typically produces non-thermal distributions. Acceleration processes are still poorly understood. A simple way to simulate the effect of particle acceleration is to inject particles at high energy. Although it does not reproduce exactly the physics involved, this prescription has been widely used and produced interesting results (as shown in most of the references cited here). Significant effort has also been taken in developing more precise modelling of acceleration mechanisms, but in such studies, the radiation field is treated crudely \\citep{LKL96,DML96,Li97,Katar06}. With the increasing number of considered processes and the increasing precision of their description, numerical analysis has become a prime method of investigation. Even so, a full treatment of the problem accounting for the coupled evolution of inhomogeneous, anisotropic distributions of leptons and photons, both in momentum and position spaces, appears to be still beyond the capabilities of present-day computers. Numerical simulations of high energy plasmas have been performed mainly following three different approaches which all make trade-offs between the various aspects of the problem. First, the Monte Carlo technique \\citep{PSS77,Stern95} allows one to follow particles and photons in space, time, and energy as they undergo mutual interactions. It solves the full radiative transfer problem and accounts explicitly for geometrical effects. At present, the MC method is probably the mos effective way to model fully 3-dimensional problems. However, this detailed procedure is time consuming, particularly when modelling the rapid dynamics of the non-thermal electron population in momentum space \\citep{mj00}, and when synchrotron self-absorption effects are important \\citep[see discussion in][]{Stern95}. For this reason, the Monte-Carlo methods have been applied to date to pure Maxwellian plasmas and/or steady state problems with 3D geometry. Another method that accounts correctly for the geometry, involves solving numerically the exact radiation transfer equation for given geometries and particle distributions \\citep{Poutanen96}. This method is far more efficient than Monte Carlo simulations which makes it easier to compare with data. It is, however, far less versatile than Monte Carlo methods and does not solve the kinetic equations for particles. The back reaction of the radiation field on the particle distribution is modelled only for the assumption of a Maxwellian plasma (in which case the plasma temperature may be adjusted according to energy balance). The method applicability is also limited to the resolution of steady state problems. The third approach, which we adopt in this paper, abandons the detailed description of the geometry to concentrate on the kinetic effects. It consists of solving the local kinetic equations for the particle and photon distributions.To maximaze efficiency, radiative transfer is usually modelled with a simple photon escape probability formalism assuming isotropic photon and particle distributions. This method can be applied to different, possibly time-dependent, problems for which geometry does not play a crucial role\\footnote{For problems in which geometry is important, radiative transfer can in principle be accounted for by coupling this kinetic code with a radiation transfer solver or a Monte Carlo code (as demonstrated by \\cite{BL01}), although computing time may then become a serious issue.}. Within the limits of the one-zone approximation, it is more efficient than other methods and allows for fast data fitting. The first detailed investigations of high energy plasmas with this technique concentrated on thermal pair plasmas \\citep{FBGPC86,Ghisellini87}. More precise modelling was then proposed in which the particle distributions were decomposed into the sum of a thermal low-energy pool and an arbitrary high energy tail \\citep{LZ87,Svensson87,Coppi92,ZLM93,GHF93,LKL96}. The latter models have been applied most to fitting and interpreting data. They do not however describe the possible deviation from a Maxwellian distribution at low energy, nor do they address explicitly any thermalization process. Only recent numerical work considered fully arbitrary distributions of particles. \\citet{GHS98} concentrated on the role of synchrotron self-absorbed radiation in AGN. They confirmed previous analytical results \\citep{GGS88} by demonstrating that the exchange of energy between particles by means of synchrotron photons can be an efficient thermalization process in magnetized sources. These simulations focused however on this specific interaction, and other processes were only considered by crude approximations, particularly Compton radiation, or not considered at all. \\citet{NM98} investigated the thermalization of arbitrary distributions by two-body particle interactions and heating by high energy protons. The additional role of synchrotron radiation was not however considered. The most complete numerical treatment of high energy plasmas was probably one developed in the context of $\\gamma$-ray bursts by \\citet{PW05}. Our code is similar to this study but differs in that these authors considered neither particle stochastic acceleration nor the effect of Coulomb losses The code presented here solves the time-dependent equations simultaneously for isotropic, arbitrary photon, electron, and positron distributions. The evolution of these populations is modelled in time while being affected by self-absorbed cyclo-synchrotron radiation, Compton scattering, pair production/annihilation, e-e and e-p Coulomb collisions, self-absorbed e-p bremsstrahlung radiation, and additional particle acceleration and heating. Each process is described with minimal approximations and by using in most of cases the exact cross sections. For instance, the formulae used for the synchrotron emission and absorption are valid from the sub-relativistic to the ultra-relativistic regime. This numerical strategy allows one to investigate many different astrophysical situations that occur in various high energy sources. The structure of this paper is as follows. Section~\\ref{sec:rad} provides a description of the microphysics adapted into our code. Then, in Sec.~\\ref{sec:numsol}, we present the numerical techniques. Finally, in Sec.~\\ref{sec:testsandapps} the code is tested against previous published results, providing an overview of its capabilities. ", "conclusions": "" }, "0808/0808.3294_arXiv.txt": { "abstract": "Results of a 3D MHD simulation of a sunspot with a photospheric size of about 20~Mm are presented. The simulation has been carried out with the MURaM code, which includes a realistic equation of state with partial ionization and radiative transfer along many ray directions. The largely relaxed state of the sunspot shows a division in a central dark umbral region with bright dots and a penumbra showing bright filaments of about $2$ to $3$ Mm length with central dark lanes. By a process similar to the formation of umbral dots, the penumbral filaments result from magneto-convection in the form of upflow plumes, which become elongated by the presence of an inclined magnetic field: the upflow is deflected in the outward direction while the magnetic field is weakened and becomes almost horizontal in the upper part of the plume near the level of optical depth unity. A dark lane forms owing to the piling up of matter near the cusp-shaped top of the rising plume that leads to an upward bulging of the surfaces of constant optical depth. The simulated penumbral structure corresponds well to the observationally inferred interlocking-comb structure of the magnetic field with Evershed outflows along dark-laned filaments with nearly horizontal magnetic field and overturning perpendicular (`twisting') motion, which are embedded in a background of stronger and less inclined field. Photospheric spectral lines are formed at the very top and somewhat above the upflow plumes, so that they do not fully sense the strong flow as well as the large field inclination and significant field strength reduction in the upper part of the plume structures. ", "introduction": "The physical understanding of sunspot structure has been hampered for decades by 1) insufficient resolution of the fine structure by observations, 2) lack of information about the layers below the visible surface, and 3) insufficient computational power to perform ab-initio 3D MHD simulation of a full sunspot including the surrounding granulation. Recently, we have seen considerable progress on all three of these fronts: 1) adaptive optics, image selection and reconstruction at ground-based telescopes and the advent of spectro-polarimetry in the visible from space with the {\\em Hinode} satellite have led to a wealth of new information about the fine structure of sunspot umbrae and penumbrae \\citep[e.g.][]{Bharti:etal:2007, Langhans:etal:2007, Ichimoto:etal:2007, Riethmueller:etal:2008, Rimmele:Marino:2006} 2) local helioseismology has started to probe the sub-surface structure of sunspots \\citep[e.g.,][]{Cameron:etal:2008}, and 3) the ever-increasing computational power of parallel computers have made ab-initio simulations of full sunspots come into reach. While \\citet{Cameron:etal:2007b} simulated solar pores of up to about 3~Mm diameter and did not find indications for the development of a penumbral structure, the first attempt to simulate a sunspot together with the surrounding granulation is due to \\citet{Heinemann:etal:2007}. They considered a rectangular section of a (slab-like) small sunspot of about 4~Mm diameter. The main result of this simulation is the formation of filamentary structures in the outer part of the spot, various properties of which (such as dark cores, outflows, and strongly inclined magnetic field) are consistent with observational results. However, the filaments found by \\citet{Heinemann:etal:2007} are much shorter than the typical lengths of real penumbral filaments and the overall extension of the simulated penumbra is very small. Here we report about results of a sunspot simulation with the {\\em MURaM} code \\citep{Voegler:etal:2005}. The simulated sunspot has a total diameter of about 20~Mm, shared about equally by umbra and penumbra. ", "conclusions": "The properties of our simulated sunspot are consistent with the general picture of sunspot structure that has emerged from observational studies. This applies to the overall structure (e.g., distinction between umbra and penumbra, average `radial' profiles of the magnetic field components and inclination angle, average outflow in the penumbra) as well as to the detailed properties of the fine structure of umbra and penumbra. The penumbral filamentation results from magneto-convective energy transport in the form of hot rising plumes, very similar to the process giving rise to umbral dots \\citep{Schuessler:Voegler:2006}. The inclined magnetic field near the periphery of the spot causes a symmetry breaking which leads to elongated filaments with strong outflows along flow tubes of nearly horizontal field near optical depth unity. In addition to the flow along the filament, the upflow also turns over into a motion perpendicular to the filament axis. Dark lanes appear above the strongest upflows owing to the upward bulging of the surface of optical depth unity and the piling up of plasma in a cusp-shaped region at the top of the filament, above which the less inclined field outside the filament becomes laterally fairly homogeneous. The horizontal outflows are concentrated along the dark lanes. All these properties are consistent with recent observational results \\citep[e.g.,][]{Bellot-Robio:etal:2005, Rimmele:Marino:2006, Langhans:etal:2007, Ichimoto:etal:2007, Borrero:etal:2008,Noort:Rouppe:2008, Zakharov:etal:2008}. Our results are also consistent with many properties of the short penumbral filaments found by \\citet{Heinemann:etal:2007}, who gave interpretations along very similar lines. The fact that simulations with two rather different numerical codes lead to basically the same picture for the formation of penumbral structure indicates that, in spite of all differences in detail, the simulations have indeed captured essential physical processes. The explanation for the Evershed effect as a natural consequence of rising plumes in an inclined field \\citep[cf.][]{Scharmer:etal:2008} connects this flow directly to the basic magneto-convective structure of the penumbra, so that it should occur whenever penumbral structure is present. The geometry of the flow pattern is such that the observable average outflow velocity need not to be connected with a net outflux of mass. This does not exclude the additional presence of siphon flows \\citep[e.g.,][]{Degenhardt:1993, Montesinos:Thomas:1997, Solanki:etal:1994}, but it is much less clear whether the pressure gradients required for a sustained outward siphon flow are maintained always and everywhere in all penumbrae. What can be said on the basis of our simulation results about the various models that have been proposed to explain the penumbral structure? First of all, we do not see evidence for the `moving flux tube' model and interchange convection \\citep[e.g.][]{Schlichenmaier:etal:1998b, Schlichenmaier:etal:1998a}. In our simulations, the progression of the filament heads toward the umbra during their formation phase is not caused by the inward motion of a narrow flux tube, but rather due to the expansion of the sheet-like upflow plumes along the filament. Furthermore, the simulations combine various aspects of the `embedded flux tube' model \\citep{Solanki:Montavon:1993, Borrero:etal:2005, Borrero:etal:2006} and the `gappy penumbra' model \\citep{Spruit:Scharmer:2006,Scharmer:Spruit:2006}. However, the simulated penumbral filaments are neither intrusions of field-free plasma from below nor are they confined to almost horizontal flux tubes disconnected from their environment, particularly in depth. The uppermost part of the plume structure forming the penumbral filament with its strong horizontal flow and almost horizontal field could possibly be represented by a kind of embedded flux tube. The main feature missing in the embedded flux tube models is the overturning convection within the filament and the deep-reaching upflow of plasma that provides the primary energy supply. In this respect, the underlying plume structure with its reduced (albeit non-vanishing) field strength has much similarity with the `gappy' configuration of the penumbra. This scenario captures the convective origin of the penumbral filamentation, even though the gaps form within the strong magnetic field. As a consequence, the gaps contain a horizontal field and, in most cases, are not connected to the almost field free convecting plasma below the penumbra. There is no clear evidence in our simulations that the penumbral structure is affected or even caused by the fluting instability as suggested by \\citet{Weiss:etal:2004}. However, the periodic boundary condition in the horizontal ($x$) direction used in the simulation implicitly corresponds to the existence of identical sunspots just about 20~Mm from the penumbral boundaries, which certainly affects the field structure, particularly the inclination, in the outer penumbra. Therefore, the simulation possibly does not well represent the convective pumping effect suggested by \\citet{Weiss:etal:2004} as a mechanism for the downward dragging of magnetic flux in the outer penumbra. This might provide a possible explanation for the still rather small extension of the simulated penumbra. Observations in fact indicate that penumbrae are often suppressed on the side of a sunspot which faces a nearby spot of the same polarity. Altogether, our results indicate a new level of realism in the theoretical modelling of sunspot structure. The properties of the simulated penumbral filaments are consistent with a variety of observational results and provide a basis for a physical understanding of umbral and penumbral structure in terms of magneto-convective processes. On the other hand, there are still clear discrepancies between the numerical results and real sunspots, so that there is some way ahead to be covered towards a completely satisfactory model. Our penumbral structure does not yet appear to be fully evolved and the overall extension of the penumbra is still somewhat small. The average intensity profile indicates that we have simulated the development of an inner penumbra, while the outer penumbra might be more strongly affected by convective pumping \\citep{Weiss:etal:2004}. The lower boundary condition remains arbitrary since we still have no reliable observational constraints concerning the subsurface structure of sunspots. Computational limits have forced us to use a rather coarse spatial resolution of 32~km and a fairly small computational box. Furthermore, we could only cover a relatively short overall evolution time. As a consequence, the effective diffusivities in the simulation are still much larger than the real values. Test calculations with different resolution show that 1) first indications for filamentary structure appear already at a horizontal resolution of 96~km and 2) the reduction of field strength in the plumes increases somewhat when we move to a resolution of 24~km. On that basis, the fundamental physical process of sheet-like plume convection appears to be a robust feature. The results will certainly change in detail (and, hopefully, become even more similar to the observed penumbrae) as resolution increases, but we do not expect totally new processes replacing those that we have described here. The rapid increase in available computational power and the foreseeable progress in local helioseismology will soon alleviate some of the limitations of the present approach and thus enable us to carry out even more realistic simulations." }, "0808/0808.3777_arXiv.txt": { "abstract": "Growth of massive black holes (MBHs) in galactic centers comes mainly from gas accretion during their QSO/AGN phases. In this paper we apply an extended So{\\l}tan argument, connecting the local MBH mass function with the time-integral of the QSO luminosity function, to the demography of MBHs and QSOs from recent optical and X-ray surveys, and obtain robust constraints on the luminosity evolution (or mass growth history) of individual QSOs (or MBHs). We find that the luminosity evolution probably involves two phases: an initial exponentially increasing phase set by the Eddington limit and a following phase in which the luminosity declines with time as a power law (with a slope of $\\sim -1.2$---$-1.3$) set by a self-similar long-term evolution of disk accretion. Neither an evolution involving only the increasing phase with a single Eddington ratio nor an exponentially declining pattern in the second phase is likely. The period of a QSO radiating at a luminosity higher than 10\\% of its peak value is about 2--3$\\times10^8\\yr$, during which the MBH obtains $\\sim 80\\%$ of its mass. The mass-to-energy conversion efficiency is $\\simeq0.16\\pm0.04 ^{+0.05}_{-0}$, with the latter error accounting for the maximum uncertainty due to Compton-thick AGNs. The expected Eddington ratios in QSOs from the constrained luminosity evolution cluster around a single value close to 0.5--1 for high-luminosity QSOs and extend to a wide range of lower values for low-luminosity ones. The Eddington ratios for high luminosity QSOs appear to conflict with those estimated from observations ($\\sim 0.25$) by using some virial mass estimators for MBHs in QSOs unless the estimators systematically over-estimate MBH masses by a factor of 2--4. We also infer the fraction of optically obscured QSOs $\\sim 60-80\\%$. The constraints obtained above are not affected significantly by MBH mergers and multiple-times of nuclear activity (e.g., triggered by multiple times of galaxy wet major mergers) in the MBH growth history. We discuss further applications of the luminosity evolution of individual QSOs to obtaining the MBH mass function at high redshifts and the cosmic evolution of triggering rates of nuclear activity. ", "introduction": "\\label{sec:intro} Massive black holes (MBHs), probably remnants of QSOs \\citep{LyndenBell}, have been detected in the nuclei of many nearby galaxies \\citep{KR95, Magorrian98, Richstone, KG01, FF05}. How do these local MBHs form and evolve, and what is the most important mechanism shaping the mass distribution of MBHs? The current consensus is that the local MBHs obtained their mass mainly through accretion during phases of nuclear activity when they appeared as QSOs/AGNs,\\footnote{Hereafter, we frequently use the term QSOs rather than QSOs/AGNs, if not otherwise specified, to represent QSOs and/or AGNs for convenience.} similar to the ones seen now in the distant universe \\citep[e.g.,][]{YT02,YL04a, Marconi04, Shankar04, Barger05, Hopkins06, Shankar07}. The evolution of mass accretion onto a MBH is equivalent to the luminosity evolution, given the mass-to-energy conversion efficiency, and is recorded in the luminosity function (LF) of QSOs. However, the QSO LF depends mainly on two functions: (1) ${\\cal G}(z;\\mbh)$, the rate of nuclear activity triggered at different redshifts $z$ for MBHs with present-day mass $\\mbh$; (2) ${\\cal L}(\\tau;\\mbh)$, the luminosity evolution history of a QSO, of which the remnant MBH has a present-day mass $\\mbh$, as a function of the age of its nuclear activity $\\tau$. One cannot derive these two functions only from the knowledge of the QSO LF without additional assumptions. In an extended version of the \\citet{S82} argument, the local MBH mass distribution function (BHMF) is related to QSOs found in the distant universe by the simple integral equation \\begin{eqnarray} \\int^{\\infty}_0 \\Psi_{L}(L,z)\\left|\\frac{dt}{dz}\\right|dz & = & \\int^{\\infty}_{0} n_{\\bh}(\\mbh,t_0) \\times \\nonumber \\\\ & & \\tau\\life(\\mbh)P(L|\\mbh) d\\mbh, \\label{eq:relation} \\end{eqnarray} where $t_0$ is the present cosmic time, $n_{\\bh}(\\mbh,t_0)$ is the local BHMF, defined so that $n_{\\bh}(\\mbh,t_0)d\\mbh$ gives the number density of local MBHs with present-day mass in the range $\\mbh\\rightarrow \\mbh+d\\mbh$, $\\Psi_{L}(L,z)$ is the QSO LF, defined so that $\\Psi_{L}(L,z)dL$ gives the comoving number density of QSOs with nuclear luminosity in the range $L\\rightarrow L+dL$ at redshift $z$, \\be \\tau\\life(\\mbh)=\\int dL\\sum_{k} \\frac{1}{\\left|\\frac{d{\\cal L}(\\tau;\\mbh)}{d\\tau}|_{\\tau=\\tau_k(L,\\mbh)}\\right|} \\label{eq:taulife} \\ee is the time interval (or the QSO lifetime) in which that a MBH with present-day mass $\\mbh$ appeared as a QSO, and $\\tau_k(L,\\mbh)$ $(k=1,2,...)$ are the roots of the equation ${\\cal L}(\\tau;\\mbh)-L=0$ (see details of the derivation in \\citealt{YL04a}). Here ${\\cal L}(\\tau;\\mbh)$ represents the luminosity of a QSO and its associated MBH with present-day mass $\\mbh$ at a time $\\tau$ after the triggering of nuclear activity. The value of $\\tau\\life$ depends on the detailed definition of ``active nuclei'' or the lower threshold set to the nuclear luminosity. Finally, \\be P(L|\\mbh)=\\frac{1}{\\tau\\life(\\mbh)}\\sum_{k} \\frac{1}{\\left|\\frac{d{\\cal L}(\\tau;\\mbh)}{d\\tau}|_{\\tau=\\tau_k(L,\\mbh)}\\right|} \\label{eq:PLMbh} \\ee is the probability distribution function of the nuclear (bolometric) luminosity $L$ over the growth history of the MBH. The right-hand-side of equation (\\ref{eq:relation}) gives the total time spent per unit $L$ at luminosity $L$ by the progenitors of all the local MBHs in a unit comoving volume, which should be the time integral of the QSO LF, i.e., the left-hand-side of the equation. Multiplying equation (\\ref{eq:relation}) by the BH mass accretion rate $(1-\\epsilon)L/(\\epsilon c^2)=\\dot\\bh$ (see eqs.~\\ref{eq:mdotinf} and \\ref{eq:mdot} below), where $\\epsilon$ is the mass-to-energy conversion efficiency and $c$ is the speed of light, and then integrating it over cosmic time $t$ reduces to the So{\\l}tan (1982) argument \\citep{YL04a}. Provided that two basic quantities, i.e., the local BHMF and the QSO LF, can be observationally determined with sufficient accuracy, the kernel $\\tau\\life(\\mbh)P(L|\\mbh)$, containing information on the luminosity evolution history of individual QSOs/MBHs, may be solved from the integral equation~(\\ref{eq:relation}). Therefore, the extended So{\\l}tan argument is expected to give robust but more detailed constraints on the growth of MBHs than the simple energetic argument due to \\citet{S82}. As an alternative approach to the theoretical models based on the hierarchical co-evolution of MBHs and galaxies/galactic halos studied intensively in the literature \\citep[e.g.,][]{ER88, HR93, HNR98, KH00, Granato01, WL03, Volonteri03, Croton06, Bower06,Malbon07}, in this paper we use the integral equation (\\ref{eq:relation}) to statistically constrain the growth history of individual MBHs or ${\\cal L}$. The advantages of this approach are: (1) the accretion history of individual QSOs, ${\\cal L}(\\tau;\\mbh)$, is isolated from the triggering rate of nuclear activity, ${\\cal G}(z;\\mbh)$, which is presumably associated with mergers of galaxies or instabilities of galactic disks; and (2) it is free of the many adjustable parameters introduced in the co-evolution models and probably also avoids uncertain assumptions on seed BHs. Note that these two functions, ${\\cal G}(z;\\mbh)$ and ${\\cal L}(\\tau;\\mbh)$, are mixed in the differential continuity equation for BHMF evolution presented in \\citet[][see also \\citealt{Cavaliere71}, \\citealt{CP89}, and \\citealt{CP90}]{SB92}, which is widely used in studying the growth of MBHs \\citep[e.g.,][]{Marconi04, Shankar07}. Using the luminosity evolution curves, i.e., ${\\cal L}(\\tau;\\mbh)$, obtained from numerical simulations of colliding galaxies, \\citet{Hopkins06} elaborated a unified model for the origin of QSOs and MBHs (see also their other papers listed therein). A possible concern with that approach is that simulations of colliding galaxies have a spatial resolution much larger than the scale of accretion disks around MBHs and therefore may not reflect the real luminosity evolution, as the disk accretion is probably self-regulated in the vicinity of MBHs rather than being directly determined by the material infall rate from a much larger scale or the Bondi-accretion rate (see discussions in \\S~\\ref{sec:models}). (For another model of the possible light curve, see \\citealt{CO07}.) Estimating the local BHMF can be done with recent advances in observations \\citep[e.g.,][]{Salucci99,AR02,YL04a,Marconi04,Shankar04,Lauer07a,Tundo07}. First, MBHs are believed to exist in the nuclei of most, if not all, nearby galaxies \\citep{KR95,Magorrian98,Richstone,KG01,FF05}. Second, it has been well established that tight correlations exist between the MBH mass and various galactic properties, such as mass, luminosity, stellar velocity-dispersion, light concentration and binding energy of the hot components of galaxies \\citep[here hot components mean either ellipticals or spiral bulges;][]{KR95, Magorrian98, FM00, Gebhardt00, Tremaine02, HR04, MH04, Graham01, AR07}. Third, the luminosity or velocity-dispersion functions of nearby galaxies have been well determined by large surveys such as the Sloan Digital Sky Survey \\citep[SDSS;][]{Blanton03, Bernardi03, Sheth03}. Combining the correlation between the MBH mass and galaxy velocity dispersion (or luminosity) with the velocity-dispersion (or luminosity) distribution of nearby galaxies, we estimate the local BHMF in \\S~\\ref{sec:BHMF}. In the past several years, the QSO LF has been determined over unprecedentedly large luminosity and redshift ranges both from optical surveys such as the Two Degree Field QSO Redshift Survey (2Qz) and SDSS, and from X-ray surveys by ASCA, Chandra and XMM-Newton. For example, the optical QSO LF has been obtained over the redshift range $0.41.4$. We find that LIRGs at $0.61.4$ within 1~Mpc and $\\pm \\, 500 \\, \\rm km \\, s^{-1}$). The contrast between the activities of the close environments of LIRGs and ULIRGs appears especially enhanced in the COSMOS field density peak at $z\\sim0.67$, because LIRGs on this peak have a larger fraction of passive neighbours, while ULIRGs have as active close environments as those outside the large-scale structure. The differential environmental activity is related to the differences in the distributions of stellar mass ratios between LIRGs/ULIRGs and their close neighbours, as well as in the general local density fields. At $0.8200\\times$ the Milky Way value) forces a large fraction of the CO gas into neutral carbon. Dust is important for H$_2$ and HD formation, already at metallicities of $10^{-4}-10^{-3}$ solar, for electron abundances below $10^{-3}$. ", "introduction": "In the study of primordial chemistry, and subsequently the formation of the first stars, it is crucial to understand the ability of interstellar gas to cool through atomic and molecular emissions and to collapse. Furthermore, atomic and molecular species allow for probes of the ambient conditions, like density and temperature, under which stars form. The basic questions of this contribution are: What sets the abundances of atoms and molecules that cool gas? What is the role of radiation, metallicity and dust in molecule formation? A number of different chemical processes are relevant to this effect: \\smallskip Ion-molecule reactions: A$^+$ + BC$\\rightarrow$AB$^+$ + C; Neutral-neutral reactions: A + BC$\\rightarrow$AB + C; Dissociative recombination: AB$^+$ + e$^-$$\\rightarrow$A + B; Radiative recombination: A$^+$ + e$^-$$\\rightarrow$A + $h\\nu$; Radiative association: A + B$\\rightarrow$AB + $h\\nu$; Ionization: A + CR/UV/X-ray$\\rightarrow$A$^+$ + e$^-$; Dissociation: AB + UV$\\rightarrow$A + B; Charge transfer: A$^+$ + B$\\rightarrow$A + B+; Grain surface reactions: Grain + A + B$\\rightarrow$Grain + AB. \\smallskip In any chemical network, the above reactions play an important role. For example, the charge transfer between H$^+$ and O, followed by reactions with H$_2$ to H$_3$O$^+$, and dissociative recombination with e$^-$, leads to species like OH and H$_2$O (following certain branching ratios). Similarly complex routes exist for CO. In any case, many species are typically joined through different chemical routes. Thus, it is not trivial to construct concise chemical networks if one wants to include important molecules like CO and H$_2$O\\footnote{It is important to realize that water can be quite an important heating agent in the presence of a warm infrared background, like $T>50$ K dust or a $z>15$ CMB (Spaans \\& Silk 2000).}. Of course, in the limit of low metallicity, chemistry simplifies. Basically, no metals implies no molecules except for H$_2$ and HD (and a few minor species). Still, even small amounts of metals and dust ($\\sim 10^{-4}$ solar) can be crucial to the efficient formation of species like H$_2$, HD, CO, H$_2$O and many others, which is the purpose of this contribution. ", "conclusions": "Radiative feedback effects differ for UV and X-ray photons at any metallicity, with molecules surviving quite well under irradiation by X-rays. Starburst and AGN will therefore enjoy quite different cooling abilities for their dense molecular gas. The presence of a cool molecular phase is strongly dependent on metallicity. Strong irradiation by cosmic rays ($>200\\times$ the Milky Way value) forces a large fraction of the CO gas into neutral carbon. Dust is important for H$_2$ and HD formation, already at metallicities of $10^{-4}-10^{-3}$ solar. Finally, one should always solve the equations of statistical equilibrium to distinguish properly between the excitation, radiation and kinetic temperature of a system. I.e., the thermodynamic floor set by the CMB is only a hard one if the density is high enough (larger than the critical density of a particular transition) to drive collisional de-excitation. \\bigskip\\bigskip\\bigskip" }, "0808/0808.0542_arXiv.txt": { "abstract": "A period of slow contraction with equation of state $w > 1$, known as an ekpyrotic phase, has been shown to flatten and smooth the universe if it begins the phase with small perturbations. In this paper, we explore how robust and powerful the ekpyrotic smoothing mechanism is by beginning with highly inhomogeneous and anisotropic initial conditions and numerically solving for the subsequent evolution of the universe. Our studies, based on a universe with gravity plus a scalar field with a negative exponential potential, show that some regions become homogeneous and isotropic while others exhibit inhomogeneous and anisotropic behavior in which the scalar field behaves like a fluid with $w=1$. We find that the ekpyrotic smoothing mechanism is robust in the sense that the ratio of the proper volume of the smooth to non-smooth region grows exponentially fast along time slices of constant mean curvature. ", "introduction": "For over two decades, the only known mechanism for homogenizing, isotropizing and flattening the universe was inflation, a period of accelerated expansion with an equation of state $w$ (ratio of pressure to energy density) near -1. Its success in resolving the horizon and flatness problems is a principal reason why inflation became an essential part of the standard model of cosmology. In recent years, an alternative mechanism has been discovered in which smoothing and flattening occurs before the big bang as the universe undergoes a period of slow contraction with $w >1$. This alternative, known as the ekpyrotic mechanism \\cite{Khoury:2001wf,Erickson:2003zm}, has been incorporated in alternatives to standard big bang inflationary cosmology including the ``ekpyrotic\" \\cite{Khoury:2001wf}, ``new ekpyrotic\" \\cite{Buchbinder:2007ad} and cyclic models \\cite{Steinhardt:2002ih}. We note that both the inflationary and ekpyrotic mechanisms can produce nearly scale-invariant spectra for density perturbations in addition to smoothing and flattening the universe. Until now, the ekpyrotic mechanism has only been shown to work in cases where the deviations from smoothness and flatness are small and perturbative when the ekpyrotic phase begins. The purpose of this paper is to show that the ekpyrotic mechanism is powerful and robust enough to smooth the universe even when the initial perturbations are large and non-linear. Let us first review how the inflation and the ekpyrotic mechanism work in the perturbative regime where the cosmic evolution is well approximated by the Friedmann equation with an anisotropy term: \\begin{equation} \\label{eq1} H^2 = \\frac{8 \\pi G}{3} \\left(\\frac{\\rho_m^0}{a^3} + \\frac{\\rho_r^0}{a^4}+ \\frac{\\rho_w^0}{a^{3(1+w)}} \\right) - \\frac{k}{a^2} + \\frac{\\sigma^2}{a^6}, \\end{equation} where $H \\equiv \\dot{a}/{a}$ is the Hubble parameter; $a(t)$ is the Friedmann-Robertson-Walker scale factor normalized so that the value today, $t_0$, is $a(t_0)=1$; $\\rho_i^0$ represents the present value of the energy density for component $i$, where $m$ represents non-relativistic matter, $r$ represents radiation and $w$ represents an energy component with equation of state $w$, such as a scalar field and its potential. The last two terms on the right-hand side represent spatial curvature and anisotropy. For an expanding universe, the term that dominates Eq.~(\\ref{eq1}) after a long period of expansion is the one with the smallest power of $a$ in the denominator. With only radiation and matter, the dominant term would be the spatial curvature, leading to a universe that is unacceptably open or closed by the present epoch. However, introducing an energy component with $w \\approx -1$ totally changes the outcome because this component ($\\rho_w$) then has the smallest exponent and dominates the Einstein equation, while the curvature and anisotropy (and other energy components) become negligible. This is the essence of how inflation works if the initial conditions are perturbative. For inflation, there are several ``Cosmic No-hair'' theorems in addition to numerical results supporting the claim that homogeneity, isotropy and flatness develop even when the initial conditions are non-linear and non-perturbative~\\cite{kolb_turner_86,Goldwirth}. Now consider the analogous arguments for a contracting universe. With $a(t)$ shrinking, the dominant term in Eq.~(1) will be the one with the largest exponent of $a$ in the denominator. For a universe with matter and radiation only, this would be the anisotropy term, which famously overtakes the evolution and drives the universe into chaotic mixmaster behavior. On the other hand, if there is an energy component with $w >+1$, then this energy component dominates instead of anisotropy or spatial curvature, and chaotic mixmaster behavior never begins \\cite{Erickson:2003zm}. For a scalar field $\\phi$ with potential energy density $V(\\phi)$, the ratio of pressure to energy density is \\begin{equation}\\label{eq2} w \\equiv \\frac{ \\frac{1}{2} \\dot{\\phi}^2 -V}{\\frac{1}{2} \\dot{\\phi}^2 +V}, \\end{equation} which can be significantly greater than unity when $V$ is less than zero and non-negligible and which approaches $w=1$ if the scalar field kinetic energy dominates. The ekpyrotic phase in ekpyrotic and cyclic models includes an effective scalar field component of this type. For the cyclic model, the ekpyrotic phase is preceded by a period of dark energy domination and accelerated expansion which, if sustained long enough, would make the universe uniform and flat before the ekpyrotic contraction phase begins. In this case, the perturbative argument above should be reliable and sufficient to conclude that the universe is smooth and flat as it approaches the big crunch. However, in the ekpyrotic or new ekpyrotic models, generally, or in the cyclic model with a very short dark energy phase, the conditions at the beginning of the ekpyrotic phase are under less control. This paper investigates the robustness of the ekpyrotic smoothing and flattening mechanism when the initial conditions are non-linear and non- perturbative to determine how the situation compares with the perturbative case and with inflation. For this study we are not concerned with any particular form for the initial conditions that may be motivated by some specific model; rather, we would like to understand how a generic, highly inhomogeneous and anisotropic space-time evolves under the influence of the proposed smoothing mechanism, modeled here by a scalar field with a negative exponential potential. Such negative exponential potentials arise naturally in supergravity and in string theory. Investigation of the proposed smoothing mechanism requires numerical solution of the coupled Einstein-scalar system of equations. For simplicity, we restrict our studies to deviations from smoothness along a single spatial dimension. We use an orthonormal frame representation of the equations written in terms of Hubble-normalized, scale invariant variables~\\cite{Uggla:2003fp} similar to that described in \\cite{Curtis:2005va}, though here coupling to a scalar field instead of a fluid, and using constant-mean-curvature (CMC) time slices. We discretize the equations using second-order accurate finite difference techniques, and solve them with a variant of the Berger and Oliger~\\cite{BO} adaptive mesh refinement (AMR) algorithm for coupled elliptic-hyperbolic equations~\\cite{Pretorius:2005ua}. We find smooth regions that are scalar field dominated in which the scalar field (kinetic plus potential energy density) component behaves like a fluid with $w \\gg 1$, and also regions where the scalar field kinetic energy dominates over the potential energy and the scalar field behaves like a fluid with $w=1$. These latter regions remain inhomogeneous and anisotropic, and throughout this paper we will refer to these parts of the universe as the ``anisotropic regions''. Note however that the anisotropic regions are neither anisotropy nor matter dominated because both the scalar field and the anisotropy of the metric play important roles in the dynamics. Futhermore, note that despite the fact that matter in the smooth regions behaves effectively like a fluid with $w \\gg 1$ there is no issue of superluminal propagation as might arise from an actual fluid with such an equation of state. This is because we are always solving the scalar wave equation with potential where disturbances always propagate within the light cone. In the anisotropic regions ``spikes'' also form, which are places where the fields change on very small spatial scales, and are similar to regions with this property that have been observerd in numerical simulations of singularities in vacuum spacetimes\\cite{dgspike}. Despite the presence of the scalar field, the anisotropic regions exhibit dynamical behavior similar to chaotic mixmaster vacuum solutions, where there are a series of relatively quick transitions between longer epochs where the solution can be described by a $w=1$ Bianchi type I spacetime. A difference here though is that there are only a {\\em finite} number of transistions, so the mixmaster behavior terminates after several transitions. These dynamics are also known to occur in spacetimes where the matter {\\em is} a fluid with $w=1$~\\cite{Curtis:2005va, Coley05}. AMR is necessary to resolve the spiky features that form both in the anisotropic regions and, in some instances, briefly in what will eventually become smooth scalar field dominated regions, and to resolve the almost domain wall-like transitions that develop between the smooth and anisotropic regions. % The outline for the rest of the paper is as follows. In Sec. \\ref{sec_methods} we describe the equations, initial conditions, and numerical methods used to solve them. We present the results in Sec. \\ref{sec_results}. The primary conclusion is that a scalar field with a potential inspired by cyclic models is a remarkably powerful and robust smoothing mechanism during a contracting phase of the universe, able to drive the spacetime to homogeneity and isostropy even starting with highly non-linear deviations from an FRW spacetime. Concluding remarks and a discussion of future work is given in Sec. \\ref{sec_conclusion}. ", "conclusions": "Our computations provide evidence that the ekpyrotic mechanism for smoothing and flattening the universe is robust and powerful, comparable qualitatively and quantitatively to the inflationary mechanism incorporated in the conventional big bang model. This evidence is the behavior as the singularity is approached of a class of spacetimes that, while not completely general, contain several degrees of freedom and begin far from FRW. Both the inflationary and the ekpyrotic mechanism require the addition of an energy component that is commonly mocked up as a scalar field with potential energy. For inflation, the important feature is that, for some initial conditions, the scalar field can act like a fluid with $w \\approx -1$. It has been shown numerically that, beginning from highly non-linear and highly irregular initial conditions, regions dominated by the scalar field and with $w \\approx -1$ come to dominate the volume of the universe\\cite{Goldwirth}. Similarly, we have found the regions in which the scalar field acts like a fluid with $w >1$ come to dominate exponentially the volume of the universe during a contracting phase. This result addresses one of the key criticisms raised when the ekpyrotic model of the universe was first introduced; namely, it was suggested that the model required smooth initial conditions \\cite{Kallosh:2001ai}. One of the motivations for extending the ekpyrotic picture into a cyclic model was to include a period of dark energy domination before the ekpyrotic phase began in order to prepare smooth conditions \\cite{Steinhardt:2002ih}. Now, from the results here, it is clear that the dark energy epoch is not required for this purpose. Not only does this allow the possibility that the dark energy phase lasts only a few e-folds in the cyclic picture, as suggested in \\cite{Erickson:2006wc}, but it also opens the way for more general bouncing cosmologies that incorporate the ekpyrotic mechanism but do not cycle. With these results in hand, we are now prepared to tackle the bounce itself in the case that it is non-singular ($a(t)$ shrinks to a non-zero value and then begins to increase). For the non-singular bounce, the equation of state must decrease from $w>1$ to $w< -1$ for a finite period during which anisotropy and inhomogeneity grows. Our goal is to determine if their growth can be kept at a level consistent with observations, establishing the viability of these bouncing cosmological models." }, "0808/0808.2292_arXiv.txt": { "abstract": "{\\psr is one of only 9 known double neutron star systems. These systems are highly valuable for measuring the masses of neutron stars, measuring the effects of gravity, and testing gravitational theories. } {We determine an improved timing solution for a mildly relativistic double neutron star system, combining data from multiple telescopes. We set better constraints on relativistic parameters and the separate masses of the system, and discuss the evolution of \\psr in the context of other double neutron star systems.} {\\psr has been regularly observed for more than 10 years by the European Pulsar Timing Array (EPTA) network using the Westerbork, Jodrell Bank, Effelsberg and Nan\\c cay radio telescopes. The data were analysed using the updated timing software \\tempo.} {We have improved the timing solution for this double neutron star system. The periastron advance has been refined and a significant detection of proper motion is presented. It is not likely that more post-Keplerian parameters, with which the individual neutron star masses and the inclination angle of the system can be determined separately, can be measured in the near future. } {Using a combination of the high-quality data sets present in the EPTA collaboration, extended with the original GBT data, we have constrained the masses in the system to $m_\\mathrm{p}<1.17$~\\msun and $m_\\mathrm{c}>1.55$~\\msun ($95.4\\% $ confidence), and the inclination angle of the orbit to be less than $47$ degrees (99\\%). From this we derive that the pulsar in this system possibly has one of the lowest neutron star masses measured to date. From evolutionary considerations it seems likely that the companion star, despite its high mass, was formed in an electron-capture supernova. } ", "introduction": "\\psr was discovered in the Green Bank Northern Sky Survey \\citep{snt97}. It is one of only 9 double neutron star systems (DNSs) known. The pulsar orbits its companion neutron star in 8.63~days, and its 40~ms spin period, together with the derived low surface magnetic field, is typical of a mildly recycled pulsar. \\cite{nst96} first described the system and have already shown that the space velocity is probably quite low. They also measured the periastron advance from which the total system mass was estimated. \\cite{tc99}, using a Bayesian analysis with no constraints on the orientation of the orbit, found the masses of the pulsar and its companion to be $m_\\mathrm{p}= 1.56^{+0.20}_{-1.20}$ \\msun and $m_\\mathrm{c}= 1.05^{+1.21}_{-0.14}$ \\msun ($95\\% $\\,confidence). An improved timing solution was presented by \\cite{hlk+04}, although not discussed in detail. DNSs are presently the best available tool for testing strong-field gravity effects. The \\psr\\ orbit is only mildly relativistic, making it of limited use for tests of gravitational theories. However, any constraints on its post-Keplerian (PK) parameters, the inclination of its orbit, or the masses of the pulsar and its companion are highly valuable for studies of the evolution of these systems (e.g. \\citealt{plp+04,kbk+07}). Like PSR~J1811$-$1736 \\citep{cks+07}, \\psr\\ has a rather wide orbit. This could indicate that the evolution into a DNS system has been slightly different than most other systems, which generally have tight orbits with periods of several hours. However, \\psr\\ does follow the recently discovered spin period-eccentricity relation \\citep{fkl+05,dpp05}, suggesting that the evolution cannot be too different from the tight DNSs. \\psr has been observed regularly using the four 100~m class radio telescopes in Europe. This has allowed us to select high-quality data resulting in the best possible analysis of the system parameters to date. For completeness, we have also included data from telescopes at Green Bank as presented in \\cite{nst96, nts99}. We present the new timing solution and discuss the limits our solution puts on the inclination and the masses of the neutron stars in the system. We discuss the future prospects of detecting multiple PK parameters, and we related our results to the evolution of DNSs. ", "conclusions": "\\subsection{Masses}\\label{section:masses} \\begin{figure} \\centering \\includegraphics[width=6.4cm, angle=270]{square.pdf} \\caption{Mass-mass diagram for \\psr and its companion. The hatched region is excluded for $i > 90$ degrees, the dash-dotted lines are indicating constraints resulting from the \\xdot\\ measurement, $i < 69$ degrees and shapiro delay limit, $i < 47$ degrees. The diagonal line constrains $\\dot\\omega$, with errors. The dotted lines for constant pulsar and companion mass indicate the range of neutron star masses measured in other DNSs \\citep{tc99,lbk+04,fkl+05}. \\label{fig:massplot}} \\end{figure} For any given theory of gravity, measuring PK parameters will put constraints on the masses and the inclination of the system, depending only on the individual masses and the Keplerian parameters (e.g. \\citealt{sta03}). Figure~\\ref{fig:massplot} shows all the current restrictions we were able to derive from our timing solution (Table~\\ref{tab:jwnge-2}). The improved, highly significant \\omdot\\ measurement gives us a very accurate determination of the total mass. From the non-detection of Shapiro delay parameters it is clear that the inclination of the orbit is quite low. Furthermore, the $\\dot{x}$ measurement presented in \\S 3.3, taken at face value, confirms that the system is at low inclination angles. The relativistic time-dilation/gravitational redshift parameter, $\\gamma$, is only measurable when the periastron advance is large enough to decouple its effect on the timing residuals from the measurement of the semimajor axis and its time derivative \\citep{bt75, mt77, dt92}. Changing $\\omega$ by only a few degrees will take several hundred years, and a detection of $\\gamma$ can therefore not be expected in the near future. Moreover, \\xdot\\ turns out to be highly covariant with $\\gamma$ in the timing analysis. This makes an independent measurement of $\\gamma$ impossible and obscures the interpretation of the apparent measurement of \\xdot. To clarify the influence of all the PK perturbations on our timing data, and to calculate accurate, refined values of the stellar masses, we performed a comprehensive, self-consistent timing analysis which simultaneously incorporated all relativistic and kinematic phenomena in the timing solution. Our primary goal was to measure or constrain the values of $m_\\mathrm{p}$ and $m_\\mathrm{c}$. This procedure also finds constraints on the two angles which describe the orientation of the orbit, $i$ and $\\Omega$. For convenience, we always represent the latter as an offset from the proper motion position angle, i.e. $\\Theta_\\mathrm{\\mu}-\\Omega$. This angle runs between $-180^\\circ$ and 180$^\\circ$, while inclination runs between $0^\\circ$ and $180^\\circ$. The four degrees of freedom in this problem ($m_\\mathrm{p}$, $m_\\mathrm{c}$, $i$, and $\\Theta_\\mathrm{\\mu}-\\Omega$) are reduced to three by noting that the two masses and the inclination are related via the Keplerian mass function equation: \\begin{equation}\\label{eqn:kepler} f\\equiv\\frac{(m_\\mathrm{c}\\sin i)^3}{(M_\\mathrm{T})^2} = \\frac{x^3}{T_\\mathrm{\\odot}}\\left(\\frac{2\\pi}{P_\\mathrm{b}}\\right)^2, \\end{equation} where all quantities on the right side of the equation are measured to high precision. We analysed a grid of timing solutions in a three-dimensional parameter space following the approach as explained in detail in \\cite{sna+02, sns+05}. For the three variables, we used $M_\\mathrm{T}$, $\\cos i$, and $\\Theta_\\mathrm{\\mu}-\\Omega$, and we assumed a uniform prior in each of these. For randomly oriented binary systems, $\\cos i$ and $\\Theta_\\mathrm{\\mu}-\\Omega$ each follow uniform distributions, and since \\omdot\\ is well known, only a small range of $M_\\mathrm{T}$ values need be considered, so a uniform prior is acceptable for this variable as well. For each point in our three dimensional grid, we used Eq.~\\ref{eqn:kepler} to calculate the value of $m_\\mathrm{c}$ and the corresponding $m_\\mathrm{p}$. We then used the masses and orientation angles to calculate the kinematic perturbation parameters (Eqs.~\\ref{eq:xdot},~\\ref{eq:omdot}) and relativistic timing parameters according to GR (e.g. \\S\\,4.1 of \\cite{sta03}). The \\omdot\\ used in the analysis was the sum of the kinematic and relativistic terms held fixed while all other pulsar timing parameters (astrometric, rotational, and Keplerian orbital parameters) were free to vary. We recorded the $\\chi^2$ of each timing solution, assigning a probability to each grid point based on the difference between its $\\chi^2$ and the global minimum. We used this ensemble of probabilities to calculate confidence regions for the parameters of interest and to place limits on the stellar masses. The results are given in Figs. \\ref{fig:abcd}, \\ref{fig:chi7} and \\ref{fig:paramplot}. Figure \\ref{fig:abcd} shows the orientations of the orbit allowed by the timing solution. The figure is a projection of the 95.4\\% confidence volume of the three dimensional analysis grid onto the two dimensional space shown. There are four regions allowed by the timing data. The Keplerian orbital parameters, combined with the lack of Shapiro delay restrict $i$, but because the Shapiro delay depends on $\\sin~i$, the resulting constraint is degenerate: if $i$\\, is allowed, so is $180^\\circ-i$. Since there is a detectable \\xdot, Eq. \\ref{eq:xdot} restricts $\\Theta_\\mathrm{\\mu}-\\Omega$ to two possibilities (one positive and one negative value) for any given value of $i$. Thus there are four regions of allowed solutions, labeled $A$, $B$, $C$ and $D$ in the figure. Figure \\ref{fig:chi7} shows 68.3\\%, 95.4\\%, and 99.7\\% confidence limits on $\\cos~i$ and $M_\\mathrm{T}$ space. For reference, the figure shows lines of constant mass difference $m_\\mathrm{c}-m_\\mathrm{p}$, at intervals of 0.2\\msun, as calculated from $i$ and $M_\\mathrm{T}$, and it indicates the region at low inclination angle that is excluded because $m_\\mathrm{p}>0$ is not satisfied in that region. The confidence limits were calculated by marginalizing over the probability values for values of $\\Theta_\\mathrm{\\mu}-\\Omega$ for a given combination of $\\cos~i$ and $M_\\mathrm{T}$. The best-fit solutions have relatively high $m_\\mathrm{c}-m_\\mathrm{p}$, i.e. high companion masses and low pulsar masses, although the confidence contours stretch to much lower values of $m_\\mathrm{c}-m_\\mathrm{p}$. The determination of $M_\\mathrm{T}$ is dominated by the relativistic \\omdot, but the obtained values are perturbed by the kinematic $\\delta$\\omdot, given by Eq. \\ref{eq:omdot}. This perturbation splits the allowed values for the total mass into two regions. This can be understood by plugging the values of $\\Theta_\\mathrm{\\mu}-\\Omega$\\, from each of the four regions of Fig.~\\ref{fig:abcd} into Eq.~2, and noting that $\\csc~i$ is always positive, so that $\\delta$\\omdot\\ is positive for solutions $A$ and $B$ and negative for solutions $C$ and $D$. This means that the observed $\\delta$\\omdot\\ has been biased towards higher values for solutions $A$ and $B$, so that the true relativistic $\\delta$\\omdot\\ is lower than that calculated without the kinematic correction. Since $M_\\mathrm{T}$ is proportional to $\\delta$\\omdot$^{3/2}$, this means that the true total mass is lower for solutions $A$ and $B$. Similarly, the true mass is higher for solutions $C$ and $D$. Figure \\ref{fig:paramplot} shows the 95.4\\% confidence volume projected into several two-dimensional parameter spaces. For this figure, the values of $i$ and $\\Theta_\\mathrm{\\mu}-\\Omega$ correspond to solution $D$, but the distributions of all other quantities are essentially identical for all four solution regions. We note that, for completeness, all possible values for the parameters are shown in Figs. \\ref{fig:abcd}, \\ref{fig:chi7} and \\ref{fig:paramplot} however we consider it unlikely that the pulsar mass will be lower than $1$\\msun. We used the probabilities from the grid analysis to constrain $m_\\mathrm{p}$ and $m_\\mathrm{c}$. The results are essentially identical for all four solution regions. The central 95.4\\% confidence intervals are $m_\\mathrm{p}=0.72^{+0.51}_{-0.58}$\\,\\msun and $m_\\mathrm{c}=2.00^{+0.58}_{-0.51}$\\,\\msun. If, instead of central confidence intervals, we use 95.4\\% confidence upper and lower limits for the individual masses, these convert to $m_\\mathrm{p}<1.17$\\,\\msun and $m_\\mathrm{c}>1.55$\\,\\msun. Note that the apparent discrepancy between these numbers originates from different areas covered in the probability distribution. For the total mass, the 95.4\\% confidence intervals are $2.7188\\pm0.0011$\\,\\msun for solutions $A$ and $B$, and $2.7217\\pm0.0018$\\,\\msun for solutions $C$ and $D$. However, allowing any of the four solutions yields the range $2.720_{-0.002}^{+0.003}$\\msun. This last number is the most accurate value for $M_\\mathrm{T}$, as it reflects the uncertainty as to which orientation of the orbit is correct. \\begin{figure} \\centering \\includegraphics[width=7.0cm, angle=270]{fig6new.pdf} \\caption{Allowed values for orbital orientation $i$ and $\\theta_\\mathrm{\\mu}-\\Omega$. The figure shows a projection of the 95.4\\% confidence volume of the three dimensional grid analysis of timing solutions onto this two dimensional space. See the text for discussion. \\label{fig:abcd}} \\end{figure} \\begin{figure} \\centering \\includegraphics[width=7.0cm, angle=270]{fig7new.pdf} \\caption{Allowed values for total mass and $\\cos~i$. The two sets of contours are 68.3\\%, 95.4\\%, and 99.7\\% confidence limits on these quantities for solutions $A$ and $B$ (lower, white contours) and $C$ and $D$ (upper, gray contours). Dotted lines indicate values of $m_\\mathrm{c}-m_\\mathrm{p}$, at intervals of 0.2\\msun. The dark region on the right of the plot is excluded as it does not satisfy $m_\\mathrm{p}>0$. \\label{fig:chi7}} \\end{figure} \\begin{figure} \\centering \\includegraphics[width=10.0cm]{fig8new.pdf} \\caption{Values of relativistic $\\gamma$, kinematic $\\delta\\dot{\\omega}$, kinematic $\\dot{x}$, orbital orientation $\\theta_\\mathrm{\\mu}-\\Omega$, and masses $m_\\mathrm{p}$ and $m_\\mathrm{c}$ allowed by the timing solution, as a function of inclination angle. The figure shows a projection of the 95.4\\% confidence volume of the three dimensional grid analysis of timing solutions onto each of the two dimensional spaces shown. The regions corresponds to solution $D$, but other solutions give essentially identical results. \\label{fig:paramplot}} \\end{figure} \\subsection{Evolution} The system is valuable for studies of DNS evolutionary scenarios. It has been believed for a long time that supernova explosions result in high kick velocities on the remaining neutron star \\citep{bai89}. Recently, it has been argued that different evolutionary scenarios can be possible in binary systems with particular mass and chemical properties \\citep{plp+04,heu07}. The so-called electron-capture collapse of an O-Ne-Mg core of a helium star leads to a {\\it fast} supernova explosion which is believed to be symmetric and therefore does not result in a kick of the second-born neutron star. For most DNSs in which proper motion measurements have been possible, a low space velocity has been derived. Our measurement of the proper motion of \\psr, implying a velocity of about 25 km\\,s$^{-1}$, is consistent with the evolutionary scenario mentioned above, and another piece of evidence that not all pulsars receive a large kick at birth. Van den Heuvel (2007) argues that the low velocity of DNS systems may be correlated with the masses of the second-formed neutron stars in DNSs being somewhat lower than normal: $1.25(6)$\\msun. Our 95.4\\% limit\\footnote{The 99.7\\% limit is $m_\\mathrm{c} > 1.39$\\msun} of $m_\\mathrm{c} > 1.55$\\msun appears to contradict this correlation. In case of a symmetric supernova the mass lost during the explosion can be calculated by the expression $e=\\Delta M_\\mathrm{SN}/M_\\mathrm{T}$ \\citep{bv91}, where $e$ is the eccentricity and $M_\\mathrm{T}$ the total mass after the supernova explosion. For \\psr\\ this results in $\\Delta M_\\mathrm{SN}=0.68$\\msun. If we use our 95.4\\% lower limit on the companion of 1.55~\\msun this leads to a core-progenitor mass of 2.23\\msun, which is considered to be in the lower range for helium cores \\citep{dp03b}, and the progenitor star must have had a mass of about $9$\\msun. As these stars are believed to produce degenerate O-Ne-Mg cores \\citep{plp+04,phlh08}, this is another indication that the companion neutron star must have formed in an electron-capture core collapse. The somewhat higher companion mass can possibly be explained by assuming extra fall-back during the supernova event of $\\sim0.2$\\msun (e.g. \\citealt{fk01,fry07}). To allow for this amount of fallback the supernova explosion has probably been very weak. An electron-capture core collapse is the most probable explanation for the parameters of this system. However, constraining the masses even better would be very interesting. \\subsection{Search for the companion} Apart from the original discovery papers \\cite{nst96,snt97}, there is no report of searches for pulsations of the companion of \\psr. The Green Bank Northern Sky Survey, optimized to find millisecond pulsars, had a flux-density limit at 370 MHz of 8 mJy for slow pulsars (with periods above 20 ms). The very high characteristic age and the low magnetic field imply that \\psr is the first-born and recycled neutron star in the system. The (unseen) second neutron star will be the young object, with a long spin period, which has probably already slowed down significantly and possibly passed the death-line to become undetectable as a radio pulsar. Moreover, if the second neutron star is still active as a radio pulsar, it has a reasonable chance of being beamed away from us. In this case, taking into account that the precession timescale of the spin axis is very long, it is not expected that the companion, if active as a pulsar, will rotate into view on a short timescale. However to be certain that no weak pulsed signal from the second neutron star has been missed, we have searched several WSRT observations of 30 minutes at 840, 1380 and 2300 MHz for pulsations of the companion of \\psr. We performed an acceleration search on the data although, because of the wide orbit, the expected smearing due to the orbital motion of the companion is likely to be negligible. In the only double pulsar system known so far, PSR~J0737$-$3039A/B, the second pulsar is only visible for a small part of each orbit \\citep{lbk+04}. To take this possibility into account, and assuming the second pulsar would be visible in an equal fraction of the orbit as PSR~J0737$-$3039B, we have searched another set of observations spread equally across the full orbital phase range. In all searches no evidence of pulsations of a companion were found to the limits of 1.1, 0.25 and 0.5 mJy at 840, 1380 and 2300 MHz respectively, which means that if the companion is active as a pulsar, and beamed towards us, it must be less luminous than 0.098 mJy~kpc$^2$ at 1380 MHz." }, "0808/0808.3227_arXiv.txt": { "abstract": "Previous solar observations have shown that coronal loops near 1\\,MK are difficult to reconcile with simple heating models. These loops have lifetimes that are long relative to a radiative cooling time, suggesting quasi-steady heating. The electron densities in these loops, however, are too high to be consistent with thermodynamic equilibrium. Models proposed to explain these properties generally rely on the existence of smaller scale filaments within the loop that are in various stages of heating and cooling. Such a framework implies that there should be a distribution of temperatures within a coronal loop. In this paper we analyze new observations from the EUV Imaging Spectrometer (EIS) on \\textit{Hinode}. EIS is capable of observing active regions over a wide range of temperatures (\\ion{Fe}{8}--\\ion{Fe}{17}) at relatively high spatial resolution (1\\arcsec). We find that most isolated coronal loops that are bright in \\ion{Fe}{12} generally have very narrow temperature distributions ($\\sigma_T \\lesssim 3\\times10^5$\\,K), but are not isothermal. We also derive volumetric filling factors in these loops of approximately 10\\%. Both results lend support to the filament models. ", "introduction": "High spatial resolution solar observations have shown that coronal loops with temperatures near 1\\,MK have properties that are difficult to reconcile with physical models. Loops at these temperatures persist for much longer than a radiative cooling time, suggesting quasi-steady heating. The densities inferred from the observations, however, are much higher than can be reproduced by steady, uniform heating models. The temperature gradients along the loops are also much smaller than predicted by the simple models. (e.g., \\citealt{lenz1999,aschwanden2000b,winebarger2003}). Several models have been proposed to explain the properties of coronal loops at these temperatures. \\cite{aschwanden2000b}, for example, suggested that the observed loops were actually composed of smaller scale threads that were steadily heated at their footpoints. Footpoint heating leads to somewhat higher densities and flatter temperature gradients relative to steady heating models. At high densities, however, loops at these temperatures can become thermodynamically unstable (e.g., \\citealt{mok2005,muller2004,winebarger2003}), leading to catastrophic cooling. Multi-thread, impulsive heating models have also been suggested (e.g., \\citealt{warren2003}). In these models the fact that loops cool much more rapidly than they drain accounts for the high densities. Multiple threads in various stages of heating and cooling are needed to explain the observed lifetimes and temperature gradients. In both cases, these multi-thread models indicate the need for emission formed over a range of temperatures to reproduce the observed intensities (see also \\citealt{reale2000}). One limitation of the observational results from \\textit{TRACE} is that they are derived from narrowband filtergrams with somewhat limited diagnostic capabilities. The launch of the EUV Imaging Spectrometer (EIS) on the \\textit{Hinode} mission provides us with an opportunity to revisit some of these observational results using spectroscopic data. EIS is a high spatial and spectral resolution spectrometer that covers much of the same wavelength range as \\textit{TRACE}. EIS has a very broad temperature coverage and can image the solar corona in individual emission lines from the lower transition region to the hottest flares. In this paper we focus on measuring the emission measure distribution in coronal loops near 1\\,MK. We have selected 20 relatively isolated loop segments from several different active region observations and computed differential emission measure distributions from the background subtracted loop intensities. For this work we focus on loops that are bright in \\ion{Fe}{12} and find that for these loops the distribution of temperatures is almost always narrow, with a dispersion of several times $10^5$\\,K. We also find volumetric filling factors of approximately 10\\%. These results support the idea that coronal loops are composed of smaller scale filaments that are below the spatial resolution of current solar instruments. ", "conclusions": "Some previous work has suggested that coronal loops, as currently observed, are isothermal. For example, \\cite{aschwanden2005b} find that the majority of narrowest loops observed with TRACE are consistent with an isothermal DEM. Since TRACE is limited to observations in only three channels (\\ion{Fe}{9}, \\ion{Fe}{12}, and \\ion{Fe}{15}) it is difficult to distinguish between an isothermal distribution and the narrow distributions that we measure spectroscopically. The general absence of \\ion{Fe}{15} emission in the loops that we have studied is consistent with \\cite{aschwanden2005b}. \\cite{delzanna2003} also found examples of relatively cool ($\\sim0.9$\\,MK) nearly isothermal loops observed with low resolution spectroscopic data. These results also suggested filling factors near 1. Our filling factor results are smaller than this, but we also find that the filling factor to be inversely proportional to the loop pressure (also see \\citealt{warren2008}). We do see some loops with a relatively broad emission measure distribution ($\\log\\sigma_T\\sim5.7$), which is consistent with the results of \\cite{schmelz2007} and \\cite{patsourakos2007}. Our sample, which is small, suggests that such loops are rare, however. These new observational results lend support to the non-equilibrium, multi-thread models of these ``warm'' coronal loops. It remains to be seen if hydrodynamic models can reproduce the observed loop properties. The combination of high densities and narrow temperature ranges will be difficult to reconcile with nanoflare models (e.g., \\citealt{patsourakos2006}). The narrow temperature distributions suggest that these filaments are evolving coherently." }, "0808/0808.3820_arXiv.txt": { "abstract": "Precession in an accretion-powered pulsar is expected to produce characteristic variations in the pulse properties. Assuming surface intensity maps with one and two hotspots, we compute theoretically the periodic modulation of the mean flux, pulse-phase residuals and fractional amplitudes of the first and second harmonic of the pulse profiles. These quantities are characterised in terms of their relative precession phase offsets. We then search for these signatures in 37 days of X-ray timing data from the accreting millisecond pulsar XTE J1814$-$338. We analyse a 12.2-d modulation observed previously and show that it is consistent with a freely precessing neutron star only if the inclination angle is $< 0.1^\\circ$, an a priori unlikely orientation. We conclude that if the observed flux variations are due to precession, our model incompletely describes the relative precession phase offsets (e.g. the surface intensity map is over-simplified). We are still able to place an upper limit on $\\epsilon$ of $3.0 \\times 10^{-9}$ independently of our model, and estimate the phase-independent tilt angle $\\theta$ to lie roughly between $5^\\circ$ and $10^\\circ$. On the other hand, if the observed flux variations are not due to precession, the detected signal serves as a firm upper limit for any underlying precession signal. We then place an upper limit on the product $\\epsilon \\cos \\theta$ of $\\leq 9.9 \\times 10^{-10}$. The first scenario translates into a maximum gravitational wave strain of $10^{-27}$ from XTE J1814$-$338 (assuming a distance of 8 kpc), and a corresponding signal-to-noise ratio of $\\leq 10^{-3}$ (for a 120 day integration time) for the advanced LIGO ground-based gravitational wave detector. ", "introduction": "\\label{intro} Accreting millisecond pulsars (AMSPs) are a subset of neutron stars in low-mass X-ray binaries (LMXBs) that exhibit persistent X-ray pulsations with periods below 10 ms. In the standard recycling scenario, AMSPs are the evolutionary link between LMXBs and nonaccreting, radio millisecond pulsars \\citep{alpar82, radha82}. Eight AMSPs have been discovered at the time of writing \\citep{wij04, morgan05, galloway07, krimm07}. Most AMSPs are X-ray transients. Once every few years, they emerge from quiescence and become detectable during an outburst lasting several weeks. The outburst is attributed to enhanced accretion [e.g. \\citet{lasota01}], funnelled onto a small number of hotspots on the star. Little is known about the shape, position, or number of these hotspots \\citep{roma04, kulk05}, but they do give rise to detectable X-ray pulsations, from which the spin period and orbital parameters can be determined. During an outburst, surface thermonuclear burning also causes type I X-ray bursts, which last a few minutes and occur on average once every few days. Type I X-ray bursts have been observed in three AMSPs to date: SAX J1808.4$-$3658, XTE J1814$-$338 \\citep{wijnands05}, and HETE J1900.1$-$2455 \\citep{vanderspek05}. In the burst tails, a small component of the X-ray flux ($\\sim 15$\\% for XTE J1814$-$338) oscillates at the spin frequency. AMSPs are expected to be relatively powerful gravitational wave sources \\citep{wattskrishnan08}. The fastest, IGR J00291+5934 \\citep{eckert04}, spins at $\\Omega_\\ast/2\\pi = 599$ Hz, well below the theoretical breakup frequency for most nuclear equations of state ($\\sim$ 1.5 kHz) \\citep{cook94, bildsten98}. Similarly, the fastest radio millisecond pulsar, PSR J1748$-$2446ad \\citep{hessels06}, and the fastest nonpulsating LMXB, 4U 1608$-$52 \\citep{hartman03}, spin at frequencies of 716 Hz and 619 Hz respectively. The gap below the breakup frequency is explained if the star is deformed by one part in $\\sim 10^{8}$, such that gravitational radiation balances the accretion torque at hectohertz frequencies \\citep{bildsten98}. Several physical mechanisms can produce the requisite deformation: magnetic mountains \\citep{pamel04, melpa05, pamel06, vigmel08}, thermocompositional mountains caused by electron capture gradients \\citep{usho00}, toroidal internal magnetic fields \\citep{cutler02}, and r-modes \\citep{anders98, owen98, nayyar06}. AMSPs are therefore promising targets for ground-based, long-baseline interferometers like the Laser Interferometer Gravitational-Wave Observatory (LIGO). An AMSP at a distance of 1 kpc, spinning at 0.4 kHz with ellipticity $\\epsilon = 10^{-8}$, generates a wave strain $h \\sim 10^{-27}$. By comparison, initial LIGO's sensitivity threshold in the 0.1--0.4 kHz band is $\\sim 10^{-26}$ during the S4 run \\citep{abbott07}. Advanced LIGO will get down to $h \\sim 10^{-27}$ in the same band, and narrowband tunability will increase its sensitivity to AMSPs further, as $\\Omega_\\ast$ is known a priori from X-ray timing. An AMSP with ellipticity $\\epsilon \\sim 10^{-8}$ is expected to precess with a period of hours to days. Magnetic mountains, for example, are built around the magnetic axis, which is misaligned in general with the rotation axis in objects which pulsate \\citep{pamel06}. More generally, a mass quadrupole of any provenance should be kicked out of alignment continuously by stochastic accretion torques \\citep{joneand02}. Hence AMSPs are promising observational candidates for observing short-period precession. Until now, however, precession has been difficult to detect in neutron stars. Only one source, the radio pulsar PSR B1828$-$11, precesses unambiguously, with period $P_{\\rm{p}} =$ 250 d \\citep{stalyshem00}. Oscillatory trends in pulse arrival times, with periods of several days, have also been reported tentatively in a few other objects \\citep{melatos00, hobbs06, pamel06}, but the physical cause is unclear. Free precession consists of a fast wobble about the angular momentum vector $\\mathbf{J}$, at approximately the pulsar spin period $P_\\ast = 2\\pi/\\Omega_{\\ast}$ and a slow retrograde rotation about the symmetry axis, with period $P_{\\rm{p}} = 2 \\pi/\\Omega_{\\rm{p}}$, which modulates the pulse shape and arrival times \\citep{zimsz79, alpines85, joneand01, joneand02, link03}. The precession frequency $\\Omega_{\\rm{p}}$ depends on the ellipticity, $\\epsilon$, and the tilt angle $\\theta$ (between the symmetry axis and \\textbf{J}), with \\begin{equation} \\label{gw} \\epsilon \\cos \\theta \\approx \\Omega_p/\\Omega_{\\star}. \\end{equation} The amplitude ratio of the gravitational wave signal at the spin frequency and its second harmonic \\citep{zimsz79, jara98}, and in the + and $\\times$ polarizations, provides \\textit{independent} information on $\\epsilon, \\theta$, the orientation of \\textbf{J}, and the emission pattern on the surface of the star. Narrowband tunability facilitates extraction of this information. In this paper, we compute theoretically the X-ray signal from a precessing pulsar for a range of orientations and compare three quantities from each pulse profile to the data: the mean flux of the profile, the zero-to-peak pulse amplitude, and the pulse-phase residuals. We search for the signature of precession in X-ray timing data from one particular AMSP, XTE J1814$-$338. An analogous search was carried out by \\citet{akgun06} for the radio pulsar PSR B1828-11, who modelled the period residuals and pulse shapes analytically taking into account precession effects (biaxial and triaxial) as well as the contribution from the magnetic spin-down torque. The authors performed searches over a range of beam locations, degrees of triaxiality, tilt angles and angle-dependent spin-down torques, finding a wide range of parameters which match the data. Thus, they were unable to constrain the shape of the star but did find that the angle-dependent spin-down torque contributes to the period residuals. Their method differs from ours in that, instead of fitting the shape of the residuals and comparing for each set of parameters, they determined the validity of a configuration by calculating Bayesian probability distribution functions for the parameters under certain constraints. The paper is structured as follows. Section \\ref{model} describes the precession model and its implementation. Sections \\ref{single} and \\ref{double} characterize the predicted X-ray signal for a biaxial, precessing pulsar with one and two hotspots respectively, specifically the relative precession phases between the flux, pulse amplitude, and pulse-phase. Section \\ref{triaxial} repeats the predictions for a triaxial, precessing pulsar. Section \\ref{datareduction} describes the data reduction and timing analysis of XTE J1814$-$338. We compare the measurements with the theory in Section \\ref{compare} and derive upper limits on $\\epsilon$, $\\theta$, and the associated gravitational wave strain in Section \\ref{conclusion}. The limit on $\\theta$ constrains the relative strength of the driving and damping forces in the system. ", "conclusions": "either the star is precessing but our surface intensity map is too simplistic, or the source is not precessing. If we attribute the 12.2-d periodicity to precession, this implies an ellipticity of $\\epsilon \\leq 3 \\times 10^{-9}$, a gravitational wave strain $h_0 \\leq 10^{-27}$, and hence a signal-to-noise ratio of $10^{-3}$ for initial LIGO and $10^{-2}$ for advanced LIGO (for a coherent 120-day search). On the other hand, if the precession is damped by internal dissipation ($\\theta$ is small), or the precession period is much longer than the 37-day data span ($\\epsilon$ is small), some other mechanism must cause the observed modulation. In this scenario, we find $\\epsilon \\cos \\theta \\leq 9.9 \\times 10^{-10}$ and $h_0 \\leq 10 \\times 10^{-27} \\cos(\\theta)^{-1}$. Although we face a negative result for this particular source, this paper establishes a framework for analyzing modulations in X-ray flux from AMSPs for a range of geometrical configurations and surface intensity maps. We anticipate that the framework will be applied to other AMSPs in the future. Given the values of $\\epsilon$ inferred from the gravitational-wave stalling hypothesis \\citep{bildsten98} and the theoretical models, e.g. of magnetic mountains \\citep{pamel06, vigmel08}, it is clear that long-term X-ray monitoring of AMSPs (over years) is essential for predicting, and then searching for, their gravitational wave signal." }, "0808/0808.0018.txt": { "abstract": "We have obtained near-infrared spectra covering the \\ion{Ca}{2} triplet lines for a large number of stars associated with 16 SMC clusters using the VLT + FORS2. These data compose the largest available sample of SMC clusters with spectroscopically derived abundances and velocities. Our clusters span a wide range of ages and provide good areal coverage of the galaxy. Cluster members are selected using a combination of their positions relative to the cluster center as well as their location in the CMD, abundances and radial velocities. We determine mean cluster velocities to typically 2.7 km s$^{-1}$ and metallicities to 0.05 dex (random errors), from an average of 6.4 members per cluster. By combining our clusters with previously published results, we compile a sample of 25 clusters on a homogenous metallicity scale and with relatively small metalliciy errors, and thereby investigate the metallicity distribution, metallicity gradient and age-metallicity relation (AMR) of the SMC cluster system. For all 25 clusters in our expanded sample, the mean metallicity [Fe/H] = $-$0.96 with $\\sigma$ = 0.19. The metallicity distribution may possibly be bimodal, with peaks at $\\sim -$0.9 dex and $-$1.15 dex. Similar to the LMC, the SMC cluster system gives no indication of a radial metallicity gradient. However, intermediate-age SMC clusters are both significantly more metal-poor and have a larger metallicity spread than their LMC counterparts. Our AMR shows evidence for 3 phases: a very early ($>11$ Gyr) phase in which the metallicity reached $\\sim -$1.2 dex, a long intermediate phase from $\\sim 10-3$ Gyr in which the metallicity only slightly increased, and a final phase from 3$-$1 Gyr ago in which the rate of enrichment was substantially faster. We find good overall agreement with the model of \\citet{pag98}, which assumes a burst of star formation at 4 Gyr. Finally, we find that the mean radial velocity of the cluster system is 148 km s$^{-1}$, with a velocity dispersion of 23.6 km s$^{-1}$ and no obvious signs of rotation amongst the clusters. Our result is similar to what has been found from a wide variety of kinematic tracers in the SMC, and shows that the SMC is best represented as a pressure supported system. ", "introduction": "The Small Magellanic Cloud (SMC) has long been recognized as being of fundamental importance for a wide variety of astrophysical studies. First, the current paradigm of galaxy formation suggests that spiral galaxy spheroids, such as the Milky Way (MW) halo, are formed by the accretion/merger of smaller, satellite galaxies (e.g., \\citealt{sea78,zen03}). As one of the nearest low mass galaxies known, the SMC is thus a possible prototype of pre-Galactic fragments. However, current $\\Lambda$CDM models suggest that the majority of the MW\u00b4s building blocks were assimilated very early in its history and that existing low mass galaxies like the SMC are survivors of this process and thus underwent a different chemical evolution (e.g., \\citealt{rob05}). At the least, many dynamical simulations (e.g., \\citealt{bek04}) suggest that the MW, Large Magellanic Cloud (LMC) and SMC compose a longterm interacting system, with the SMC and LMC likely to be eventually consumed by the MW. In addition, close encounters in the last few Gyr may well have stimulated star formation on a global scale in both the SMC and LMC (\\citealt{mur80,gar94,yos03,bek04}). These early simulations predict bursts of star formation $\\sim 0.2$Gyr ago that led to the formation of the eastern wing of the SMC and the Magellanic Bridge, and a similar close encounter event some 4 Gyr ago. Note however that the accurate proper motions crucial for properly modeling the orbits of the SMC and LMC are still problematic, and that the most recent values all suggest the Magellanic Cloud may be unbound to each other and only beginning their first close encounter with the MW (\\citealt{kal06,pia08}). The SMC, with its low global metallicity, is the best local counterpart to the host of distant dwarf irregular and blue compact dwarf galaxies, making it an attractive target for exploring the importance of metallicity in a number of contexts, including star formation, initial mass function, stellar and galaxy evolution, etc. In particular, the star clusters of the SMC, because they are (at least to first order) simple stellar populations (SSPs), are an invaluable resource with which we can explore the structure, kinematics, star formation and chemical evolution history of the SMC. As a majority of stars may have formed in clusters (see e.g., \\citealt{lad03} for the solar neighborhood and \\citealt{cha06} for the SMC, but see \\citealt{bas09} for an opposing view), their study has become even more relevant. On a cosmological scale, star clusters in the SMC are of utmost importance for the study and understanding of stellar populations in distant galaxies, since they cover age and abundance space that is not occupied by their Galactic counterparts. In addition, while the LMC cluster system suffers from the well known age gap, where only one cluster is known to have formed between $\\sim$3 and $\\sim$13 Gyr ago, (i.e.~during 3/4 of its life; \\citealt{dac91}, \\citealt{gei97}), the SMC posesses clusters that have been forming more or less continuously over the past $\\sim$11 Gyr (e.g., \\citealt{gl08b}). The SMC is the only dwarf galaxy in the Local Group that has formed and preserved populous star clusters, without a significant age gap, across (most of) the age of the Universe, and the production has been prolific; \\citet{hod86} estimates the SMC has some 2000 clusters. Despite their utility and many clear advantages, SMC clusters have been surprisingly underexploited. The number of clusters with well-determined ages from deep main-sequence photometry is minimal. Only two clusters have had their detailed chemical abundances derived via high resolution spectroscopy (NGC\\,330 - \\citealt{gon99}, \\citealt{hil99}, NGC\\,121 - \\citealt{joh04}), and an additional six clusters have metallicities that were derived from \\ion{Ca}{2} triplet (CaT) spectroscopy of individual stars (\\citealt{dch98}, hereafter DH98). Aside from these eight clusters, all other existing abundance determinations are based only on photometry or integrated spectroscopy. Thus, the really detailed information, viz. accurate ages and abundances, necessary to fully utilize SMC clusters, both as tracers of the SMC's formation and chemical evolution history, and as templates for studying stellar populations in more distant galaxies, is sorely lacking. Our group has been working to ameliorate this situation over the past several years. Using Washington photometry, we have derived ages and metallicities for almost 50 previously unstudied or poorly studied clusters \\citep{pia01,pi05a,pi07a,pi07b,pi07c}. These results have greatly improved our understanding of the global properties of the SMC cluster system, provided constraints on the age-metallicity relation (AMR), explored possible gradients, cluster formation scenarios, etc. However, these photometric cluster ages and abundances suffer from two problems. First, our photometry comes from data obtained with a 1m class telescope, which is barely adequate to reach the main sequence turn off for clusters older than several Gyr at the distance of the SMC. Second, while the photometry for the red giant stars that are used to derive cluster abundances is generally good, the Washington technique is known to require a significant age correction to metallicities derived for intermediate-age ($\\lesssim$ 5 Gyr) objects \\citep{gei03}, an age range that includes the majority of the SMC clusters so far observed. This is of course illustrative of the infamous age-metallicity degeneracy and is not a problem restricted to the Washington system; while some age and metallicity estimates from isochrone fitting to CMDs constructed in other photometric systems exist, the degeneracy between age and metallicity also makes these estimates inherently uncertain in the absence of more solid metallicity measurements based on spectroscopic data. As mentioned above, spectroscopic based abundances only exist for a handful of SMC clusters, with most of the sample coming from the work by DH98. They combined their CaT based metallicities with ages from the literature to create the first accurate AMR for the SMC, and found that it was consistent with a simple closed box model and did not require the significant gas infall or strong galactic winds that were needed to explain previous SMC AMR's \\citep[e.g.,][]{dop91}. In addition, DH98 found that their cluster velocities, like other kinematic tracers in the SMC, show no evidence for any systematic rotation of the SMC. However, their small sample size and large age errors severely limit how well we can constrain the kinematics and chemical evolution of the SMC. While the CaT method does not measure Fe abundances directly, previous authors have shown that the strength of the CaT lines are an excellent proxy for [Fe/H], and can be used in stellar populations covering a wide range of ages and abundances \\citep[e.g.,][]{col04,car07}. An added benefit of using the CaT is that it is very efficient; not only are RGB stars near their brightest in the infrared, but the multiplexing capabilities of many moderate-resolution spectrographs allows the observation of dozens of stars simultaneously, greatly increasing the probability of identifying cluster members. In a previous paper \\citep[hereafter G06]{gro06}, we have applied this method to the LMC with excellent results, following up on the pioneering work by \\citet{ols91}. Using FORS2 on the VLT, we were able to identify more than 200 member stars in 28 populous LMC clusters and determine accurate mean cluster velocities ($\\sigma$ = 1.6 km s$^{-1}$) and metallicities ($\\sigma$ = 0.04 dex) and use these to explore the global cluster metallicity distribution, kinematics, etc. To further refine our knowledge of the velocities and abundances of clusters in the SMC, we here apply this powerful technique to the SMC, again using FORS2 on the VLT to obtain medium-resolution near-infrared spectra of the CaT lines in 270 individual RGB stars in and around 16 SMC clusters. Herein we present our efforts to identify cluster members and derive mean abundances and velocities for these clusters. In a future paper we will analyze the several hundred field stars that were also observed as part of this project. This paper is organized as follows: In \\S 2, we describe our target selection process, and our spectroscopic observations and reduction procedures are detailed in \\S 3. In \\S 4 and \\S 5, we present the radial velocities and equivalent width measurements and the metallicity derivation, respectively. The membership selection process is described in \\S 6 and \\S 7 compares our metallicity values with previous determinations. In \\S 7 we also discuss our metallicity results and in \\S 8 the kinematics. Finally, in \\S 9 we summarize our results. ", "conclusions": "Magellanic Cloud clusters are an excellent laboratory for helping to unlock the secrets of cluster and galaxy formation. They are also crucial testbeds for stellar and chemical evolution models and interpreting the integrated light of distant galaxies. However, despite their utility and proximity, Small Magellanic Cloud clusters in particular have been overlooked in this regard. In order to help remedy this situation, we have carried out a large-scale investigation of the kinematics and metallicities for a number of SMC star clusters. Building on our experience with deriving these parameters for LMC clusters using the CaT technique (G06), we have used the same technique to explore SMC clusters. We obtained CaT spectra for a number of stars associated with 16 SMC clusters using the VLT + FORS2 MXU instrument. This provides the largest sample of SMC clusters with velocities and well-determined spectroscopic metallicities currently available, more than doubling the only previous study (DH98). Target clusters were selected from our Washington photometric system studies, which provide a rough estimate of age and abundance, to be old enough to possess red giant branch stars and cover a wide area across the galaxy. We used the same reduction and analysis techniques as we did in G06 to measure stellar radial velocities and metallicity. Typical radial velocities errors are 7.5 km s$^{-1}$ and metallicity errors are 0.17 dex per star. Cluster members are selected using an analysis combining their location in the CMD, position in the cluster, and radial velocities and metallicity with respect to other stars in the same field. We determine mean cluster velocities to typically 2.7 km s$^{-1}$ and metallicities to 0.05 dex (random error), from a mean of 6.4 members per cluster. A comparison of our mean cluster metallicities with those derived from Washington photometry for 11 clusters in common shows very good overall agreement, with the CaT metallicities being much more precise. This indicates that the Washington metallicities, and especially their procedure used to correct the age dependence of their metallicities for their age dependence, is appropriate. The metallicity distribution (MD), metallicity gradient and age-metallicity relation (AMR) are investigated, combining our clusters with those observed by DH98 and Glatt et al. (also with CaT) and the one cluster with a detailed, high resolution metallicity to compile a sample of 25 clusters on a homogenous metallicity scale with relatively small errors, although the ages are somewhat heterogeneous. The mean metallicity is [Fe/H] $=-0.96$, with $\\sigma = 0.19$. Dividing the sample into two age groups at 3 Gyr, the 12 older clusters have a mean metallicity of $-1.08, \\sigma = 0.17$, while the 13 younger have $-0.85, 0.15$. Most clusters lie between [Fe/H] = $-$0.75 and $-$1.25. There is a suggestion for bimodality in the MD, with peaks at [Fe/H] $\\sim -0.9$ and $-$1.15. No clear gradient is seen with distance from the center of the SMC. However, intermediate-age SMC clusters are both significantly more metal-poor and have a larger metallicity spread than their LMC counterparts. The AMR shows evidence for 3 phases: a very early ($>11$ Gyr) phase in which the metallicity reached $\\sim -1.2$, a long intermediate phase from $\\sim 10-3$ Gyr in which the metallicity only slightly increased although a number of clusters formed, and a final phase from 3-1 Gyr ago in which the rate of enrichment was substantially faster. These salient features agree with those found by most other AMR studies using other tracers and techniques. We find good overall agreement with the model of PT98 which assumes a burst of star formation at 4 Gyr. A hybrid infall $+$ outflow model of \\citet{car05} also fits the data reasonably well. The simple closed box model of DH98 yields a much poorer fit, and the AMRs derived by \\citet{har04} and \\citet{idi07} are significantly offset to higher metallicities for intermediate-age clusters. A number of different lines of evidence point to the likelihood of a burst in the SMC star and cluster formation about 3 Gyr ago. The cause of such a burst is currently a source of much speculation. The suggestion by \\citet{bek04} that it is due to a close passage of the SMC and LMC is intriguing but requires better knowledge of their orbits, especially proper motions, to be definitively tested. We finally examine the kinematics of our CaT clusters. Their mean radial velocity is = 148 km s$^{-1}$, with a velocity dispersion of 23.6 km s$^{-1}$. These values are in very good agreement with those found by DH98 from their smaller sample. Combining the 2 cluster samples, we find the kinematics are dominated by the velocity dispersion, as found in virtually all other kinematic studies of a wide variety of SMC populations.\\\\" }, "0808/0808.0189_arXiv.txt": { "abstract": "Exploring the diversity of dark energy dynamics, we discover a calibration relation, a uniform stretching of the amplitude of the equation of state time variation with scale factor. This defines homogeneous families of dark energy physics. The calibration factor has a close relation to the standard time variation parameter $w_a$, and we show that the new, calibrated $w_a$ describes observables, i.e.\\ distance and Hubble parameter as a function of redshift, typically to an accuracy level of $10^{-3}$. We discuss implications for figures of merit for dark energy science programs. ", "introduction": "} Understanding the nature of the dark energy accelerating the cosmic expansion is one of the premier questions in physics. The answer offers the possibility of deep insights into the nature of spacetime and gravity, extra dimensions, the quantum vacuum, and possibly the unification of gravitation and quantum physics. Precision mapping of the expansion history provides one path to characterizing the dark energy, in particular its equation of state and time variation. Guidance from theory is useful to predict observable signatures for cosmological probes such as distance and Hubble parameter measurements, in particular what level of accuracy is required to distinguish between models. From a model one can predict distance-redshift relations etc.\\ but the number of models is vast; one would like to identify model independent or at least generic characteristics of the dark energy. Indeed, such properties exist, as discussed in detail recently by \\cite{cahndl}, for classes of behavior in the early time evolution of dark energy, valid for $z\\gtrsim2$ when the dark energy does not strongly affect the background expansion. In this article we seek to extend characterization of the dark energy properties in terms of the equation of state to the entire observable history. This requires a different approach, calibrating the evolution through a ``stretch'' relation between the amplitude of the time variation and the time variable or scale factor of the expansion. The calibration then provides a physical basis for a compact and highly accurate parametrization of the dark energy influence on observables. In \\S\\ref{sec:dyn} we examine several diverse models, looking for similarities and distinctions. We introduce the calibration in \\S\\ref{sec:stretch} and discuss its relation to a standard parametrization of the equation of state. \\S\\ref{sec:obs} examines the utility of the description and shows that it achieves robustness and accuracy at the $10^{-3}$ level, sufficient for next generation data. We discuss some implications for figures of merit of dark energy science programs in \\S\\ref{sec:fom}. Those readers wanting to get right to the results could start in the middle of \\S\\ref{sec:stretch}. ", "conclusions": "} Having investigated a diverse group of dark energy models to explain the acceleration of the cosmic expansion, we find a homogeneous ``stretch'' relation that calibrates the time variation behavior into tight families. This stretch factor is closely related to the standard time variation measure $w_a$, and we verify that the equation of state form $w(a)=w_0+w_a(1-a)$, with $w_a$ now treated as a fit parameter to observables, delivers fractional accuracy at the $10^{-3}$ level. Such accuracy is sufficient for next generation data and the $w_0$-$w_a$ form can be viewed as an appropriate compression of the expansion history information that can be extracted from such observations. That is, this form neither overcompresses (loses important information) nor undercompresses (lacks additional leverage). This indicates there is no need nor generic benefit for going to a third parameter. Note that \\cite{barnardchi} saw similar compression and tight relations within a principal component analysis relying on many modes. To gain insight into the nature of dark energy, particular combinations of $w_0$-$w_a$ may have enhanced leverage and hence merit, separating the cosmological constant from the thawing class, each from the freezing class, and possibly zeroing in on specific models within a class. The calibration, and its robustness and accuracy in accounting for the observable relations, offers a well-defined method for assessing the next generation dark energy science program. Interpretation of those observations should offer promising insights into the physics of the accelerating universe." }, "0808/0808.2529_arXiv.txt": { "abstract": "We present $H$-band (1.65$\\mu$m) surface photometry of 57 galaxies drawn from the Local Sphere of Influence (LSI) with distances of less than 10\\,Mpc from the Milky Way. The images with a typical surface brightness limit 4 mag fainter than 2MASS ($24.5$\\,mag arcsec${}^{-2}< \\mu_{lim}<26$\\,mag arcsec${}^{-2}$ ) have been obtained with IRIS2 on the 3.9~m Anglo-Australian Telescope. A total of 22 galaxies that remained previously undetected in the near-IR and potentially could have been genuinely young galaxies were found to have an old stellar population with a star density $1-2$ magnitudes below the 2MASS detection threshold. The cleaned near-IR images reveal the morphology and extent of many of the galaxies for the first time. For all program galaxies, we derive radial luminosity profiles, ellipticities, and position angles, together with global parameters such as total magnitude, mean effective surface brightness and half-light radius. Our results show that 2MASS underestimates the total magnitude of galaxies with $\\langle\\mu_H\\rangle_{eff}$ between $18-21$\\,mag arcsec${}^{-2}$ by up to 2.5\\,mag. The S\\'ersic parameters best describing the observed surface brightness profiles are also presented. Adopting accurate galaxy distances and a $H$-band mass-to-light ratio of $\\Upsilon_{\\ast}^H=1.0\\pm 0.4$, the LSI galaxies are found to cover a stellar mass range of $5.6<\\log_{10}(\\mathcal{M}_{\\rm stars})<11.1$. The results are discussed along with previously obtained optical data. Our sample of low luminosity galaxies is found to follow closely the optical-infrared $B$ versus $H$ luminosity relation defined by brighter galaxies with a slope of $1.14 \\pm 0.02$ and scatter of $0.3$\\,magnitudes. Finally we analyse the luminosity -- surface brightness relation to determine an empirical mass-to-light ratio of $\\Upsilon_{\\ast}^H=0.78\\pm0.08$ for late-type galaxies in the $H$-band. ", "introduction": "The observational properties of nearby galaxies such as fluxes, colours, morphologies and sizes reflect their underlying physical properties (stellar/baryonic and dark matter content, star formation rates, formation history and angular momenta). Exactly how these observational and physical properties are related is still poorly understood. By technical necessity, the observational quantities are mainly based on the optical $B$-band ($390 - 480$\\,nm). However galaxies evolving in low density environments with little external stimulation for star formation often contain significant quantities of dust (eg. see \\citealt{driver07}) which can attenuate and distort their optical light profiles. In contrast, dust attenuation is vastly reduced at near-IR wavelengths and hence the near-IR provides a spectral regime where a more accurate, unaltered representation of a galaxy's underlying stellar distribution can be obtained \\citep{gavazzi96}. Furthermore, the stellar mass of most galaxies is dominated by the quiescent old stellar component whose energy output peaks at near-IR wavelengths. Even in the extreme case of Blue Compact Dwarf (BCD) galaxies, previously thought to be primeval galaxies forming their first stars at the present epoch \\citep{thuan97}, the analysis of their resolved stellar populations has revealed the presence of stars at least a few Gyrs of age (\\citealt[eg. the BCD galaxies: VII Zw 403, Mrk 178 and I Zw 36 as discussed by][respectively]{schulteladbeck98,schulteladbeck00, schulteladbeck01}; SBS 1415+437 discussed by \\citealt{aloisi05}; I Zw 18 by \\citealt{aloisi07}; and CGCG 269-049 by \\citealt{corbin08}). In order to obtain a deeper understanding of the connection between the light and matter distribution in galaxies, a representative sample of nearby stellar systems needs to be studied in detail. The Local Sphere of Influence (LSI, $D< 10\\,$Mpc) contains large numbers of early (dE) and late-type (dIrr) dwarf galaxies that make up about 85\\% of the local galaxy population~\\citep{kraan79, schmidt92, karachentsev04}. Dwarf galaxies contribute about 4\\% to the local luminosity density and about 10-15\\% to the local H\\,{\\sc i} mass density \\citep{karachentsev04}. Due to their proximity to the Milky Way, LSI galaxies are ideal for a near-IR study which includes significant numbers of dwarf systems. Previous near-IR surveys include the Two Micron All Sky Survey \\citep[2MASS,][]{skrutskie06} as well as deeper targeted galaxy surveys \\citep{gavazzi96b, gavazzi96a, gavazzi00, boselli00}. 2MASS photometry for galaxies suffers from a number of important drawbacks that are becoming more evident as the samples of independently investigated galaxies become larger. The short integration time of 2MASS observations resulted in most of the low surface brightness (LSB) dwarfs in the LSI remaining undetected, and if they were detected, 2MASS underestimated the fluxes by as much as 70\\% \\citep{andreon02}. The targeted $H$-band observations of \\cite{gavazzi96b, gavazzi96a, gavazzi00} and \\cite{boselli00} were inherently deeper however the samples included few LSB dwarfs. This serious limitation demands a deeper and higher resolution study to investigate those galaxies that were beyond the reach of photometric near-IR studies to date. A reference atlas of images needs to have the necessary spatial resolution to probe the morphological fine-structure of these nearby galaxies and contain a significant number of dwarf galaxies that are generally overlooked. A LSI sample has the additional advantage that an increasingly large number of nearby galaxies have accurately known distances. \\cite{karachentsev06} report that 214 out of 451 LSI galaxies have distance estimates (with less than 10\\% uncertainty) by means of the tip magnitude of the red giant branch (TRGB), the Tully-Fisher relation, and the surface brightness fluctuations (SBF) method (see for example \\citealt{jerjen98, jerjen01} and \\citealt{karachentsev04}). The remaining galaxies have rough distance estimates from the luminosity of their brightest stars, radial velocities or their suspected membership to a known galaxy group. The purpose of this paper is to present a near-IR $H$-band (1.65$\\mu$m) atlas of 57 LSI galaxies, probing to flux levels approximately 4 mag arcsec${}^{-2}$ or 40 times fainter than 2MASS. The majority of the galaxies presented here are much fainter than those in previous targeted surveys. We derive photometric parameters for each object such as the total magnitude, the effective radius and effective surface brightness, S\\'ersic fitting parameters, etc. Using the best distances currently available in the literature allows us to derive physical parameters such as their luminosities and stellar masses. The paper is organised as follows: we describe the sample selection in \\S \\ref{s:sample}. In \\S\\ref{s:obs} and \\S\\ref{s:photometry} we discuss the observing strategies, the data reduction, and the photometric calibration of the images. The 11 galaxies in the sample which remained undetected at our faint detection limit or had images which could not be usefully analysed are discussed in \\S\\ref{s:nondetect}. The new data is compared to 2MASS photometry and optical ($B$-band) data in \\S\\ref{s:results} and the luminosity - surface brightness relation discussed. Interesting properties of individual galaxies are described in \\S\\ref{s:interesting}. Finally, the results are summarised in \\S\\ref{s:summary}. ~\\linebreak ", "conclusions": "\\label{s:summary} We have presented the deepest $H$-band images available to date for 57 galaxies in the Local Sphere of Influence ($D<10$\\,Mpc), obtained using the near-IR camera IRIS2 at the 3.9m Anglo-Australian Telescope. Of the 68 targets, 11 remained undetected or could not be usefully analysed due to contamination by foreground stars. The surface brightness limit reaches down to $\\mu_{lim}<26$\\,mag\\,arcsec, 4 magnitudes fainter than 2MASS. The images, cleaned from Galactic foreground contamination, reveal the morphology and extent of many of the galaxies for the first time. For 56 galaxies, we derive radial luminosity profiles, ellipticities, and position angles, together with global parameters such as total magnitude, mean effective surface brightness, half-light radius, S\\'ersic parameters, and stellar mass. No genuine young galaxies have been found in this survey. Some sample galaxies were previously identified on $B$-band photographic plates but remain undetected in the near-IR. In each case there is a plausible alternative explanation for the non-detection: \\begin{itemize} \\item AM0717-571: DSS $B_J$-band morphology resembles that of a Galactic nebula, but true nature still remains unclear. \\item HIZOAJ1616-55 and SJK98 J1616-55: possibly one or two high velocity clouds. \\item KK2000-03: Superimposed star hampers analysis however the marginal detection in the H-band suggests an unusual blue galaxy. \\item KK2000-04: Originally assumed to be a companion of NGC1313 however possibly a photographic plate flaw. \\item KK2000-06: Originally assumed to be a companion of NGC1313. More likely a background galaxy at $\\approx$2250\\,km\\,s$^{-1}$. \\item NGC2784 DW1: intrinsic extreme low surface brightness dwarf satellite of NGC2784. \\end{itemize} We also detected a double nucleus in KKS2000-09 and propose to reclassify this system as a peculiar galaxy. KKS2000-25 was shown to have distinct spiral arms in the $H$-band and thus should be classified as ``Sb\". Morphology and angular size strongly suggest that this is a background galaxy beyond 10\\,Mpc. We found compelling evidence that the short integration time of 2MASS resulted in serious underestimation of a galaxy's luminosity. The magnitudes of galaxies, with $H$-band surface brightnesses fainter than 18\\,mag\\,arcsec${}^{-2}$, obtained in our study are up to 2.5\\,mag brighter than those obtained by 2MASS. As the mean effective surface brightness correlates with the luminosity of a galaxy, we expect serious selection biases for a 2MASS-based $H$-band galaxy luminosity function fainter than M$_{H}=-20$\\,mag. There is a tight correlation (correlation coefficient = 0.97) between the $B$- and $H$-band magnitudes of a galaxy and this correlation has been demonstrated over a range of 15 magnitudes. The linear transformation between the $B$- and $H$-bands has a small scatter (0.3 mag) for bright galaxies. In the dwarf regime, there is a marginal increase in scatter and possibly a slight trend for galaxies to be redder (by approximately 1 magnitude) than indicated by the transformation found for bright galaxies. The galaxy luminosity -- mean effective surface brightness relation has been analysed to derive a semi-empirical stellar mass-to-light ratio of $\\Upsilon_{\\ast}^H=0.78\\pm0.08$ in the $H$-band. All raw and reduced $H$-band images of the 57 program galaxies in this near-IR survey will be made publicly available and can be obtained via email request." }, "0808/0808.3005_arXiv.txt": { "abstract": "This paper presents an analytic perturbation approach to the dynamics of a classical spinning particle, according to the Mathisson-Papapetrou-Dixon (MPD) equations of motion, with a direct application to circular motion around a Kerr black hole. The formalism is established in terms of a power series expansion with respect to the particle's spin magnitude, where the particle's kinematic and dynamical degrees are expressed in a completely general form that can be constructed to infinite order in the expansion parameter. It is further shown that the particle's squared mass and spin magnitude can shift due to a classical analogue of radiative corrections that arise from spin-curvature coupling. Explicit expressions are determined for the case of circular motion near the event horizon a Kerr black hole, where the mass and spin shift contributions are dependent on the initial conditions of the particle's spin orientation. A preliminary analysis of the stability properties of the orbital motion in the Kerr background due to spin-curvature interactions is explored and briefly discussed. ", "introduction": "\\label{sec:1} One of the earliest and on-going research interests in general relativity concerns the dynamics of extended bodies in the presence of strong gravitational backgrounds. Considering that virtually all astrophysical objects in the Universe, such as black holes, neutron stars, and other isolated massive bodies, have at least some spin angular momentum in their formation, it is not difficult to surmise that an in-depth study of moving relativistic systems with spin is a useful endeavour. A relevant example concerns the motion of rapidly rotating neutron stars in circular orbit around supermassive black holes like ones believed to exist in the centre of galaxies, which serve as candidate sources for emitting low-frequency gravitational wave radiation that may be detected by the space-based LISA gravitational wave observatory \\cite{LISA}. A first attempt to understand the dynamics of extended bodies in curved space-time was put forward by Mathisson \\cite{Mathisson}, who showed the existence of an interaction term involving the direct coupling of particle spin to the Riemann curvature tensor generated by a background source. Steady progress was made since this first attempt, with a notable contribution made several years afterwards by Papapetrou \\cite{Papapetrou}, who proposed that the spinning particle exists within a space-time world tube containing its centre-of-mass worldline, where its associated matter field has compact support. In addition, multipole moment contributions, i.e. beyond the mass monopole and spin dipole, to the extended objects full equations of motion were considered by Tulczyjew \\cite{Tulczyjew} and others, ultimately leading to the expressions obtained by Dixon \\cite{Dixon1,Dixon2}, with a self-consistent description for all multipole moment contributions to infinite order. While the various theories of extended body motion in curved space-time differ with respect to the higher-order multipole moments, all of them recover the ``pole-dipole approximation'' identified initially by Mathisson and Papapetrou, which are satisfactory for most practical calculations, so long as the dimensions of the spinning body are small when compared to the background space-time's local radius of curvature. These truncated expressions of the full equations of motion are commonly known as the Mathisson-Papapetrou-Dixon (MPD) equations. There has been widespread interest in applying the MPD equations to the dynamics of classical spinning particles in orbit around rotating black holes, as described by the Kerr metric \\cite{Mashhoon1,Wald,Tod,Semerak,Suzuki1}. In many ways, the Kerr background is an ideal testing ground for the MPD equations, since both mass sources are spinning, which introduce interesting spin-curvature effects that impact upon the orbiting particle's overall evolution. Furthermore, it lends itself well to numerical simulations of deterministic chaos under extreme conditions \\cite{Suzuki1,Suzuki2,Hartl1,Hartl2}, as well as studies of gravitational wave generation \\cite{Mino,Tanaka} arising from spin-induced deviations away from geodesic motion. More formal study of the MPD equations have also occurred in various forms \\cite{Ehlers,Bailey,Noonan}, including a recent perturbative approach developed by Chicone, Mashhoon, and Punsly (CMP) \\cite{Chicone}, with application to the study of rotating plasma clumps propagating in astrophysical jets directed along a Kerr black hole's axis of symmetry. Another application of the CMP approximation by Mashhoon and Singh \\cite{Mashhoon2} determined analytic expressions for leading-order spin-curvature perturbations of a spinning particle's circular orbit around a Kerr black hole. This analysis is successful in reproducing the spinning particle's kinematic behaviour compared to numerical simulations of the full MPD equations for situations where, for spin magnitude $s$ and mass $m$, the M{\\o}ller radius \\cite{Mashhoon2,Moller} for the spinning particle is $s/m \\lesssim 10^{-3} \\, r$, and $r$ is the particle's radial distance away from the background mass source. However, this approximation starts to break down when $s/(m r) \\sim 10^{-2}-10^{-1}$ for $r = 10 \\, M$, where $M$ is the Kerr black hole mass, suggesting that higher-order spin-curvature coupling terms are required to more completely describe the orbital motion. It was for this initial purpose that a generalization of the CMP approximation was very recently introduced by Singh \\cite{Singh0} to incorporate higher-order analytic contributions to the perturbation approach for the MPD equations. This generalization has several nice features. For example, as a power series expansion with respect to the particle's spin magnitude, it can be extended to formally {\\em infinite order} in the expansion. In addition, it leads to expressions that are background independent, and is fully applicable to {\\em arbitrary motion} of the particle, without recourse to any space-time symmetries within the metric. As a result, this generalization is very robust, with applicability for many distinct scenarios in theoretical astrophysics, such as the modelling of globular clusters and other many-body dynamical systems in curved space-time, and also spinning particle interactions with gravitational waves, the results of which can be compared with existing treatments \\cite{Nieto,Mohseni,Kessari}. Furthermore, this approach identifies the existence of a classical analogue for ``radiative corrections'' that shift the particle's overall squared mass and spin magnitude due to higher-order spin-curvature contributions, a feature not thought about before. It would, therefore, be very useful to investigate the computational capacity of this generalization when applied to circular motion in the Kerr background. This is especially so in extreme conditions where a transition from stable to chaotic motion may be analytically identified, for comparison with existing approaches \\cite{Suzuki1,Suzuki2,Hartl1,Hartl2} which use primarily numerical methods. The purpose of this paper is to present the generalized form of the CMP approximation for the MPD equations within the context of circular motion around a Kerr black hole, and explore the derived physical consequences. It begins with Sec.~\\ref{sec:2}, which displays the full MPD equations, followed by a presentation of the formalism behind the generalized CMP approximation in Sec.~\\ref{sec:3}. Afterwards, Sec.~\\ref{sec:4} presents the formal application of the generalized CMP approximation to the case of circular motion around a Kerr black hole, up to second-order in the perturbation expansion parameter. This is followed, in Sec.~\\ref{sec:5}, by analysis of the predicted kinematic and dynamical properties of the perturbed system, including the predicted effective squared mass and spin magnitude of the spinning particle. A general discussion of the main results obtained in this paper is found in Sec.~\\ref{sec:6}, with a brief conclusion thereafter. The metric convention adopted is $+2$ signature with Riemann and Ricci tensor definitions following MTW \\cite{MTW}, and geometric units of $G = c =1$ are assumed throughout. ", "conclusions": "\\label{sec:6} This paper outlines the generalization of an analytic perturbation approach to the Mathisson-Papapetrou-Dixon equations for a spinning point particle, first introduced by Chicone, Mashhoon, and Punsly, with an application to circular motion around a Kerr black hole. The formalism shows the existence of ``radiative corrections'' to the particle's squared mass and spin magnitudes due to spin-curvature interactions, represented in power series expansion form. In performing the analysis, it is possible to semi-analytically identify the emergence of instabilities during the particle's orbital motion, which serves as a basis for a more precise treatment in the future. One of the underlying goals of the formalism presented in this paper is to determine the perturbed orbit of the spinning particle according to the generalized CMP approximation, following the approach taken earlier \\cite{Mashhoon2}. However, to do this properly requires a modification of the equations of motion to incorporate dissipative effects due to gravitational radiation, which have not yet been taken into account. Such a modification would most certainly require evaluation of the Teukolsky equations for determining the radiation effects corresponding to an adiabatic inspiral for the spinning particle's orbit. This is a non-trivial exercise with both conceptual and technical challenges to still overcome. Once this is better understood, a determination of the perturbed orbit due to spin-curvature interactions will be considered in a future publication. For now, a second paper on the generalized CMP approximation in the Vaidya background is forthcoming \\cite{Singh2} as a companion piece to accompany and compare with this paper." }, "0808/0808.1236_arXiv.txt": { "abstract": "We discuss the anisotropic arrival directions of the ultra high energy cosmic rays detected by Auger which I consider one of the biggest discoverie in astrophysics during the last year. ", "introduction": "As you may conclude from the different concluding remarks at this meeting there are various approaches to doing them. Most of my esteemed colleagues spoke on several topic that they consider important now and in the future. I decided to concentrate on one single topic that excited not only me but a large number of astrophysicists that were never interested in ultra high energy cosmic rays (UHECR) - the observed correlation of the highest energy events detected by the Pierre Auger Observatory in Argentina with active galactic nuclei (AGN)~\\cite{Auger1,Auger2}. This discovery marked the beginning of a new type of astronomy - cosmic ray astronomy - that may become as important in the future as the TeV gamma ray astronomy has recently become. TeV gamma ray astronomy does not only reveal the type of the gamma ray sources, it also provides an indirect measurement of the infrared/optical background whose estimates are based on theoretical calculations involving the light emission of different types of stars and galaxies and the cosmological evolution of these objects and the matter density around them in the Universe. Cosmic ray astronomy also involves some general features of the nearby Universe that are otherwise extremely difficult to measure such as the local (within 200 Mpc) distribution of matter and different powerful astrophysical objects, the average strength of the intergalactic magnetic fields (that could even be mapped when the sources are confirmed) and the galactic ones. I will discuss these results and report on the small contribution that I have made to this topic. It is not connected to the type of the UHECR sources, but includes a better than 3$\\sigma$ confirmation of the anisotropy detected by Auger~\\cite{Stanev08}. ", "conclusions": "" }, "0808/0808.3887_arXiv.txt": { "abstract": "We review the properties of the very young ($\\sim 2$\\,Myr) open cluster NGC\\,6383. The cluster is dominated by the massive binary HD\\,159176 (O7\\,V + O7\\,V). The distance to NGC\\,6383 is consistently found to be $1.3 \\pm 0.1$\\,kpc and the average reddening is determined to be $E(\\bv) = 0.32 \\pm 0.02$. Several pre-main sequence candidates have been identified using different criteria relying on the detection of emission lines, infrared excesses, photometric variability and X-ray emission. ", "introduction": "NGC\\,6383 ($\\alpha_{2000} = 17^h34^m48^s$, $\\delta_{2000} = -32^{\\circ}34\\farcm0$; $l_{II} = 355.69^{\\circ}$, $b_{II} = +0.04^{\\circ}$) is a rather compact open cluster which could be part of the Sgr\\,OB1 association together with NGC\\,6530 and NGC\\,6531. The cluster was originally discovered by John Herschel in 1834, and listed as h\\,3689 in Herschel's Cape Catalogue published in 1847. Most probably due to a clerical error, it also appears in Dreyer's New General Catalogue as NGC\\,6374, an identifier that is however not commonly used. Trumpler \\cite{Trumpler} pointed out that NGC\\,6383 belongs to a category of clusters in which a few dozen faint stars are closely grouped around a bright hot central star that frequently happens to be a binary system. In the case of NGC\\,6383, the cluster is centered on the O-type binary HD\\,159176 (m$_V$ = 5.7, see Fig.\\,\\ref{optical}) that dominates the emission from the cluster over a broad range of energies from the near-infrared to the X-ray domain. Apart from HD\\,159176, the cluster currently harbours no stars earlier than B1, although de Wit et al.\\ \\cite{deWit} suggested that the eclipsing binary candidate HD\\,158186 (O9.5\\,V, Marchenko et al.\\ 1998) might have been ejected from NGC\\,6383 through dynamical interactions in the cluster core. NGC\\,6383 and more specifically HD\\,159176 are likely responsible for the ionization of the H\\,{\\sc ii} region RCW\\,132 (Rodgers, Campbell \\& Whiteoak 1960) also known as S\\,11 (Sharpless 1953)\\footnote{Note that the S\\,11 identifier is not to be confused with the number of the nebula in the revised catalogue of H\\,{\\sc ii} regions published by Sharpless \\cite{Sharp2}. In the latter catalogue, the nebula corresponds to entry number 12 (in the SIMBAD database, this identifier is referred to as Sh 2-012).} or Stromlo\\,67 (Gum 1955). Rodgers et al.\\ \\cite{RCW} described RCW\\,132 as a 110\\,arcmin $\\times$ 80\\,arcmin medium brightness crescent-shaped region. Images of this emission nebula can be found for instance on plate 131 of Lyng\\aa\\ \\& Hansson \\cite{LH}. During a survey of the galactic ridge at 1390\\,MHz, Westerhout \\cite{Westerhout} detected a large ($1.5^{\\circ} \\times 1.5^{\\circ}$) ring-like radio structure roughly centered on NGC\\,6383. However, Westerhout cautioned that this feature was difficult to separate from the background emission. Part of this source is likely due to RCW\\,132. \\begin{figure}[htb] \\centering \\includegraphics[draft=False,width=\\textwidth]{ngc6383b.eps} \\caption{The region around the core of NGC\\,6383 as seen on the UKSTU survey red plate. The bright star in the middle of the field of view is the massive binary system HD\\,159176. The field size is 10\\,arcmin $\\times$ 10\\,arcmin.\\label{optical}} \\end{figure} ", "conclusions": "" }, "0808/0808.4053_arXiv.txt": { "abstract": "{} {We report the $\\gamma$-ray activity from the intermediate BL Lac S5 0716+714 during observations acquired by the AGILE satellite in September and October 2007. These detections of activity were contemporaneous with a period of intense optical activity, which was monitored by GASP--WEBT. This simultaneous optical and $\\gamma$-ray coverage allows us to study in detail the light curves, time lags, $\\gamma$-ray photon spectrum, and Spectral Energy Distributions (SEDs) during different states of activity.} {AGILE observed the source with its two co-aligned imagers, the Gamma-Ray Imaging Detector (GRID) and the hard X-ray imager (Super-AGILE), which are sensitive to the 30~MeV--50~GeV and 18--60 keV energy ranges, respectively. Observations were completed in two different periods, the first between 2007 September 4 -- 23, and the second between 2007 October 24 -- November 1.} {Over the period 2007 September 7 -- 12, AGILE detected $\\gamma$-ray emission from the source at a significance level of 9.6-$\\sigma$ with an average flux (E$>$100~MeV) of $(97 \\pm 15) \\times 10^{-8}$ photons cm$^{-2}$ s$^{-1}$, which increased by a factor of at least four within three days. No emission was detected by Super-AGILE for the energy range 18--60~keV to a 3-$\\sigma$ upper limit of 10 mCrab in 335 ksec. In October 2007, AGILE repointed toward S5 0716+714 following an intense optical flare, measuring an average flux of $(47 \\pm 11) \\times 10^{-8}$ photons cm$^{-2}$ s$^{-1}$ at a significance level of 6.0-$\\sigma$. } { The $\\gamma$-ray flux of S5 0716+714 detected by AGILE is the highest ever detected for this blazar and one of the most intense $\\gamma$-ray fluxes detected from a BL Lac object. The SED of mid-September appears to be consistent with the synchrotron self-Compton (SSC) emission model, but only by including two SSC components of different variabilities.} ", "introduction": "\\label{introduction} The source S5 0716+714 was discovered in 1979 as the optical counterpart to an extragalactic radio source (K\\\"uhr et al. 1981). Two years later, it was classified as a BL Lac (Biermann et al. 1981) because of its featureless optical spectrum and high linear polarization. The optical continuum was so featureless that every attempt to determine its spectroscopic redshift has failed. However, by optical imaging of the underlying galaxy, Nilsson et al. (2008) derived a redshift of z = 0.31 $\\pm$ 0.08. According to its spectral energy distribution the source belongs to the intermediate BL Lac class, observations by BeppoSAX (Tagliaferri et al. 2003, Giommi et al. 1999) and XMM-$Newton$ (Foschini et al. 2006, Ferrero et al. 2007) provide evidence for a concave X-ray spectrum in the 0.1--10~keV band, which is a signature of the presence of both the steep tail of the synchrotron emission and the flat part of the Inverse Compton spectrum. The detection in the X-ray band of fast variability only in the soft X-ray component can be interpreted as a slowly variable Compton component and a fast, erratic variable tail of the synchrotron component. In general, the variability of this blazar is strong in every band on both long and short (intraday) timescales. The optical and radio historical behaviour was analyzed by Raiteri et al. (2003), while the EGRET telescope on the Compton Gamma-Ray Observatory (Hartman et al. 1999) detected S5 0716+714 several times in the $\\gamma$-rays (Lin et al. 1995; von Montigny et al. 1995). The integrated flux above 100 MeV varied between (13 $\\pm$ 5) and (53 $\\pm$ 13) x 10$^{-8}$ photons cm$^{-2}$ s$^{-1}$. In this Letter, we present the analysis of the AGILE data obtained during the S5 0716+714 observations in the period 2007 September -- October, in particular two flaring episodes: the first in mid-September, the other on 2007 October 22-23. Preliminary results were communicated in Giuliani et al. (2007). The intense $\\gamma$-ray flare detected by AGILE in mid-September triggered observations by the GLAST-AGILE Support Program (GASP) of the WEBT\\footnote[1]{\\texttt{http://www.oato.inaf.it/blazars/webt/}\\\\ see e.g. Villata et al. (2006, 2007); Raiteri et al. (2006, 2007).} (see Carosati et al., 2007). About one month later, the GASP observed a bright phase of the source, triggering new AGILE and Swift observations. In the period from September to October 2007, S5 0716+714 showed intense activity with strong optical flaring episodes and a rare contemporaneous optical-radio outburst (Villata et al. 2008). The results of a multiwavelength campaign on S5 0716+714 with simultaneous Swift and AGILE observations in October 2007 are discussed in Giommi et al. (2008). Throughout this paper, the quoted uncertainties are given at the 1--$\\sigma$ significance level, unless otherwise stated. ", "conclusions": "\\label{0716:discussion} To analyze the gamma-optical correlation, we have applied the Discrete Correlation Function (DCF; see Edelson $\\&$ Krolick (1988) and Hufnagel $\\&$ Bregman (1992)) to the $\\gamma$-ray and $R$-band light curves. The DCF is a statistical method developed to analyze unevenly sampled data sets. The $R$-band flux densities were averaged over 0.1 day bins to smooth the intranight variability. The result is shown in Fig. 3. The DCF displays a significant peak (DCF $\\sim$ 0.9) for a time-lag of -1 day. Notwithstanding the large uncertainty due to poor $\\gamma$-ray sampling, this result suggests a possible delay in the $\\gamma$-ray flux variations with respect to optical variations of the order of 1 day. The uncertainty in the delay can be estimated by Monte Carlo simulations based on the ``flux randomization / random subset selection'' method (see Peterson et al. (2001) and Raiteri et al. (2003)). By performing 2000 simulations we derived a 1-$\\sigma$ uncertainty level in the lag of 1.1 days. \\begin{figure}[!t] % \\centering \\includegraphics[angle=0,scale=0.40]{fig4.eps} \\caption[0716 DCFsep]{Discrete correlation function (DCF) between the $\\gamma$-ray and $R$-band light curves for S5 0716+714 in September-October 2007. \\label{0716:fig:dcs}} \\end{figure} In Fig. 1, it is clear that most of the DCF signal originates from the quasi-simultaneity of the $\\gamma$-ray and optical peaks of late October (JD $\\sim$ 2454396-397). As for the September AGILE detection, the strong $\\gamma$-ray flare lacks strictly simultaneous optical observations since it occurred at both the start of the GASP operation and the optical observing season. \\begin{table}[!b] \\begin{center} \\caption{Parameters for the two SSC components.} \\begin{tabular}{|l|lll|} \\hline &1$^{st}$ SSC comp &2$^{nd}$ SSC comp &Units \\\\ \\hline $\\delta$ &14 &25 & \\\\ $\\Gamma$ &7.5 &15 & \\\\ $R$ &40 &40 & [$10^{15}$ cm] \\\\ $B$ &1 &0.5 & [G] \\\\ $\\gamma_{\\rm min}$ &200 &3 x 10$^{3}$ & \\\\ $\\gamma_{\\rm break}$ &4 x 10$^{3}$ &6 x 10$^{3}$ & \\\\ $p_{\\rm low}$ &2.0 &2.0 & \\\\ $p_{\\rm high}$ &4.8 &4.8 & \\\\ $n_{\\rm e}$ &2.2 &0.8 & [cm$^{-3}$] \\\\ $\\theta$ & 2 & 2 & [deg] \\\\ \\hline \\end{tabular} \\end{center} \\end{table} We note that when the $\\gamma$-ray fluxes are $\\la$ 120 $\\times$ 10$^{-8}$\\, photons cm$^{-2}$ s$^{-1}$, the corresponding optical flux densities are around 25--30 mJy. In contrast, the October $\\gamma$-ray peak reaching $\\sim$ 200 $\\times$ 10$^{-8}$\\, photons cm$^{-2}$ s$^{-1}$ has an optical counterpart of 40--45 mJy (see Fig. 1). This suggests that a significant optical event occurred at the same time as the $\\gamma$-ray flare and in September was missed. Moreover, while the ratio between the high and low $\\gamma$-ray flux levels is about 2.5, in the optical band the same ratio is of the order of 1.5. Hence, the gamma variability appears to depend on the square of changes in optical flux density. This would favour a SSC interpretation, in which the emission at the synchrotron and IC peaks is produced by the same electron population, which self-scatters the synchrotron photons. The 1-day time-lag in the high-frequency peak emission found from the DCF could then be due to the light travel time of the synchrotron seed photons that scatter the energetic electrons. The Spectral Energy Distribution for the AGILE and GASP-WEBT data of September 2007 is shown in Fig.~\\ref{0716:fig:spectrum} as green dots. The blue dashed line shows a simple SSC model that fits simultaneous observations of a ground state (see Tagliaferri et al. 2003 and references therein) and non-simultaneous EGRET data (empty blue circles). Because the high state of mid-September 2007 cannot be fitted by a one-zone SSC component alone, we used a model with two SSC components. To the first SSC component that reproduces the ground state, we add a second SSC component that dominates the optical and $\\gamma$-ray bands. Both the components are reproduced with a double power-law electron distribution: the spectral index is $p_{low}$ from $\\gamma_{min}$ to $\\gamma_{break}$ and $p_{high}$ above $\\gamma_{break}$. The parameters of the two SSC components are reported in detail in Table 1. We cannot exclude a second component due to an external seed photon field, for example mirrored by a putative broad line region, which could also account for the possible 1 day time lag. Nevertheless, the large amplitude of $\\gamma$-ray variability with respect to that of the optical favors a SSC explanation. The luminosities observed in the optical and $\\gamma$-ray ranges are both $10^{48}$ erg s$^{-1}$, with a dissipated power in the jet rest-frame of $2 \\times 10^{43} (\\delta/15)^{-4}$ erg s$^{-1}$. This output is significantly large with respect to other BL Lacs, and we estimate a global power transported into the jet $L_{tot}>3 \\times 10^{45}$ erg s$^{-1}$. This may exceed the maximum power generated by a spinning black hole of mass $10^{9}$ M$_\\odot$ in most widely known models (see e.g. Cavaliere $\\&$ D'Elia 2002). \\begin{figure}[!h] % \\centering \\includegraphics[angle=0,scale=0.08]{fig5.eps} \\caption[SED of S5~0716+714.]{ The SED of S5 0716+714, including GASP-WEBT optical data quasi-simultaneous with a AGILE-GRID gamma-ray observation in September (green dots). Historical data over the entire electromagnetic spectrum relative to a ground state of the source and EGRET non-simultaneous data are represented with blue dots. Red dots represent historical data simultaneous with a high X-ray state. \\label{0716:fig:spectrum}} \\end{figure}" }, "0808/0808.3673_arXiv.txt": { "abstract": "{We report the $\\gamma$-ray activity from the Intermediate BL Lac S5 0716+714 during 2007 September--October observations by the AGILE satellite, coincident with a period of intense optical activity of the source monitored by GASP--WEBT. AGILE observed the source with its two co-aligned imagers, the Gamma-Ray Imaging Detector (GRID) and the hard X-ray imager (Super-AGILE) sensitive in the energy range 30~MeV--50~GeV and 18--60 keV respectively, in two different periods: the first between 4 and 23 September 2007, the second between 24 October and 1 November 2007. Over the period 7--12 September, AGILE detected $\\gamma$-ray emission from the source at a significance level of 9.6-$\\sigma$ with an average flux (E$>$100~MeV) of $(97 \\pm 15) \\times 10^{-8}$ photons cm$^{-2}$ s$^{-1}$, increasing by a factor of at least four within three days. No emission was detected by Super-AGILE in the energy range 18--60~keV, with a 3-$\\sigma$ upper limit of 10 mCrab in 335 ksec. The $\\gamma$-ray flux of S5 0716+714 detected by AGILE is the highest ever detected for this blazar and one of the most intense $\\gamma$-ray fluxes detected from a BL Lac object. The Spectral Energy Distribution (SED) of mid-September seems to be consistent with the synchrotron self-Compton (SSC) emission model, but only by including two SSC components with different variability. In October 2007 AGILE repointed toward S5 0716+714 following an intense optical flare, measuring an average flux of $(47 \\pm 11) \\times 10^{-8}$ photons cm$^{-2}$ s$^{-1}$ at a significance level of 6.0-$\\sigma$. The $\\gamma$-ray flux during both AGILE pointings appears to be highly variable on timescales of 1 day.} \\FullConference{Workshop on Blazar Variability across the Electromagnetic Spectrum\\\\ April 22-25 2008\\\\ Palaiseau, France} \\begin{document} ", "introduction": "The source S5 0716+714 was classified by Biermann (1981) as a BL Lac object, because of its featureless optical spectrum and high linear polarization. The optical continuum is so featureless that every attempt to determine the spectroscopic redshift of the source has failed; however, very recently through optical imaging of the underlying galaxy was estimated a redshift of z = 0.31 $\\pm$ 0.08 (Nilsson et al. 2008). The source belongs to the Intermediate BL Lac class according to its Spectral Energy Distribution. In fact, observations by BeppoSAX (Tagliaferri et al. 2003) and XMM-$Newton$ (Foschini et al. 2006, Ferrero et al. 2007) provide evidence for a concave X-ray spectrum in the 0.1--10~keV band, a signature of the presence of both the steep tail of the synchrotron emission and the flat part of the Inverse Compton spectrum. The detection in the X-ray band of fast variability only in the soft X-ray component can be interpreted as the contemporary presence of a slowly variable Compton component and a fast and erratic variable tail of the synchrotron component. In general, the variability of this blazar is strong in every band on both long and short intraday timescales. The optical and radio historical behaviour has been analyzed by Raiteri et al. (2003), while the EGRET telescope onboard $CGRO$ (Hartman et al. 1999) detected S5 0716+714 several times in the $\\gamma$-rays (Lin et al. 1995). The integrated flux above 100 MeV varied between (13 $\\pm$ 5) and (53 $\\pm$ 13) x 10$^{-8}$ photons cm$^{-2}$ s$^{-1}$. We present the analysis of the AGILE data obtained during the S5 0716+714 observations in September--October 2007, in particular two flaring episodes: the first in mid-September, the other on 22--23 October 2007. Preliminary results were communicated in Giuliani et al. (2007) and a more detailed analysis is presented in Chen et al. (2008). The strong $\\gamma$-ray flare detected by AGILE in mid-September triggered observations by the GLAST-AGILE Support Program (GASP) of the WEBT\\footnote[1]{\\texttt{http://www.oato.inaf.it/blazars/webt/; see e.g. Villata et al. (2007)}} (see Carosati et al., 2007). About one month later the GASP observed a new very bright phase of the source, triggering Swift as well as new AGILE observations. In the period from September to October 2007, S5 0716+714 showed intense activity with strong optical flaring episodes and a rare contemporaneous optical-radio outburst (Villata et al. 2008). The results of a multiwavelength campaign on S5 0716+714 with simultaneous AGILE and Swift observations in October 2007 are discussed in Giommi et al. (2008). Throughout this paper the quoted uncertainties are given at the 1--$\\sigma$ level, unless otherwise stated. ", "conclusions": "To analyze the gamma-optical correlation we applied the Discrete Correlation Function (DCF; see Edelson $\\&$ Krolick 1988; Hufnagel and Bregman 1992; Peterson 2001) to the $\\gamma$-ray and $R$-band light curves. The $R$-band flux densities were averaged over 0.1 day bins to smooth the intranight variability. The DCF is a statistical method that was developed to analyze unevenly sampled data trains. The DCF displays a significant peak (DCF $\\sim$ 0.9) at a lag of -1 day (Fig. 2, right panel). Despite the large uncertainty due to poor $\\gamma$-ray sampling, this result suggests a possible delay of the $\\gamma$-ray flux variations with respect to the optical ones on the order of 1 day. Looking at Fig. 1, one can see that most of the DCF signal comes from the quasi-simultaneity of the $\\gamma$-ray and optical peaks of late October (JD $\\sim$ 2454396-397). We notice that when the $\\gamma$-ray fluxes are $\\leq$ 120 $\\times$ 10$^{-8}$\\, photons cm$^{-2}$ s$^{-1}$, the corresponding optical flux densities are around 25--30 mJy. In contrast, the October $\\gamma$-ray peak reaching $\\sim$ 200 $\\times$ 10$^{-8}$\\, photons cm$^{-2}$ s$^{-1}$ has an optical counterpart of 40--45 mJy (see Fig. 1). This suggests that a strong optical event simultaneous to the $\\gamma$-ray flare was missed in September, since it occured at the beginning of the optical observing season as well as the start of the GASP activity. The gamma variability seems to depend on the optical flux density changes roughly quadratically and this would favour a SSC interpretation, in which the emission at the synchrotron and IC peaks is produced by the same electron population, which self-scatters the synchrotron photons. In this case, the 1-day time lag in the high-frequency peak emission found from the DCF could be due to the light travel time of the synchrotron seed photons which scatter the energetic electrons. \\begin{figure}[!t] % \\centering \\includegraphics[angle=0,scale=0.083]{SED_pos2.eps} \\caption{ The SED of S5 0716+714 including GASP-WEBT optical data quasi-simultaneous with AGILE-GRID $\\gamma$-ray observation in September 2007 (green dots). Historical data over the entire electromagnetic spectrum relative to a ground state of the source together with EGRET non-simultaneous data is represented with blue triangles. Red triangles represent historical data simultaneous with a high X-ray state.} \\label{figure3} \\end{figure} The Spectral Energy Distribution with the AGILE and GASP-WEBT data of September 2007 is shown in Figure~\\ref{figure3} as green dots. The blue solid line shows a simple SSC model fitting simultaneous observations of a ground state (see Tagliaferri et al. 2003 and references therein) together with non-simultaneous EGRET data (empty blue triangles). Because the high state of mid-September 2007 cannot be fitted by a one-zone SSC component alone, we used a model with two SSC components. Without simultaneous X-ray data the spectrum is poorly constrained, then we show two models: one with a high hard X-ray state (red dashed line) and one with a low hard X-ray state (green solid line). The first SSC component dominates in the optical and X-ray bands and it is reproduced with a double power law electron distribution: the spectral index is $p_{low}=2$ from $\\gamma_{min}$ to $\\gamma_{break}$ and $p_{high}=4.5$ above $\\gamma_{break}$. For the high X-ray state model $\\gamma_{min}$ = 500, while the low X-ray state model has $\\gamma_{min}$ = 700; in both cases $\\gamma_{break}$ = 10$^3$. The density at the spectral break is $n_{e}=40\\,cm^{-3}$, the blob radius $R=2 \\times 10^{16}$ cm and the magnetic field $B=3$ Gauss. The second SSC component contributes primarily to the gamma range of the SED and, to a lesser extent, to the optical emission. The electron distribution is a single power law with $p=4.5$, $\\gamma_{min}=4 \\times 10^{3}$ and $n_{e}=50\\,cm^{-3}$. The blob radius is $R=10^{16}$ cm and the magnetic field $B=1.3$ Gauss. Both blobs are moving with bulk Lorentz factor $\\Gamma=15$, at an angle of $3^{\\circ}$ with respect the line of sight. We cannot exclude a second component due to an external seed photon field (e.g. mirrored by a putative broad line region) which could also account for the possible 1-day time lag, but the large amplitude of $\\gamma$-ray variability with respect to that of the optical one favours a SSC explanation." }, "0808/0808.4115_arXiv.txt": { "abstract": "Although it is well recognized that gamma-ray burst (GRB) afterglows are obscured and reddened by dust in their host galaxies, the wavelength-dependence and quantity of dust extinction are still poorly known. Current studies on this mostly rely on fitting the afterglow spectral energy distributions (SEDs) with template extinction models. The inferred extinction (both quantity and wavelength-dependence) and dust-to-gas ratios are often in disagreement with that obtained from dust depletion and X-ray spectroscopy studies. We argue that this discrepancy could result from the prior assumption of a template extinction law. We propose an analytical formula to approximate the GRB host extinction law. With the template extinction laws self-contained, and the capability of revealing extinction laws differing from the conventional ones, it is shown that this is a powerful approach in modeling the afterglow SEDs to derive GRB host extinction. ", "introduction": "} In addition to the Galactic foreground extinction, GRBs and their afterglows are subject to extinction caused by the dust within their host galaxies. Evidence for this includes --- \\begin{itemize} \\item ``{\\it Dark bursts}'' -- an appreciable fraction of GRBs with X-ray and/or radio afterglows lack an optical afterglow (Jakobsson et al.\\ 2004).\\footnote{% Prior to the launch of {\\it Swift}, nearly $\\simali$60\\% of the X-ray afterglows reportedly had no optical counterparts. Despite rapid and deep searches in the {\\it Swift} era, it was found that $\\simali$1/3 GRBs with bright X-ray afterglows remain undetected at optical wavelengths (Fiore et al.\\ 2007, Schady et al.\\ 2007). } A natural explanation for dark bursts is that they lie behind significant obscuring dust columns in their host galaxies which effectively suppresses the optical light [although some dark bursts may be intrinsically faint or occur at high redshifts (say, $z\\gtsim 5$) where the Ly$\\alpha$ break has moved through the optical bands, leading to absorption of the optical light by the Ly$\\alpha$ forest]. Indeed, Schady et al.\\ (2007) found that the X-ray afterglows of GRBs not detected by UVOT were more affected by extinction than those of GRBs with detected UVOT counterparts. The recent detection of the near infrared (IR) afterglows of some GRBs (which would have been considered as ``dark bursts'' since their afterglows were not detected in any bluer bands) provides another piece of evidence for dust obscuration (e.g. see Jaunsen et al.\\ 2008, Tanvir et al.\\ 2008). \\item {\\it Reddening} -- some GRB afterglows with low redshifts appear very red, due to effects of extinction -- ultraviolet (UV)/visible light is extinguished more by dust than red light (e.g. see Klose et al.\\ 2000, Levan et al.\\ 2006). Dust reddening is also indicated by the significant deviation of the optical/near-IR spectral energy distributions (SEDs) of many afterglows from that expected from standard models. Also because of dust reddening, the Balmer line ratios in the spectra of some GRB host galaxies (e.g. see Djorgovski et al.\\ 1998), known as the {\\it Balmer decrement}, deviate from the expected ratios for the standard Case B recombination, which are fairly independent of physical conditions (Osterbrock \\& Ferland 2006). \\item {\\it Depletion} -- dust-forming heavy elements such as Si and Fe were found to be substantially depleted from the gas phase in some host galaxies (e.g. see Savaglio et al.\\ 2003). This indirectly shows the presence of dust in GRB host galaxies since the missing heavy elements must have been locked up in dust grains. \\item {\\it Connection between long GRBs and massive stars} -- there are multiple strong lines of evidence that long-duration ($\\simgt 2\\s$) GRBs are associated with the death of massive stars, occurring in regions of active star formation embedded in dense clouds of dust and gas (see Woosley \\& Bloom 2006). \\end{itemize} A precise knowledge of the extinction (quantity, wavelength-dependence) and the nature (size, composition, and quantity) of the dust in GRB host galaxies is crucial for \\begin{itemize} \\item Correcting for the extinction of afterglows from X-ray to near-IR wavelengths to derive their intrinsic luminosities -- this is particularly important for studying the luminosity distribution of GRB afterglows and their intrinsic SEDs (e.g. see Kann et al.\\ 2008); \\item Constraining the nature of the GRB progenitors (i.e. collapsing massive stars or merging neutron stars) -- if long-duration GRBs are indeed linked to the collapse of massive stars, it is most likely that their optical and near-IR afterglows will suffer from significant attenuation in the star-forming molecular clouds heavily enshrouded by dust -- the birth place of these short-lived ($\\simali 10^{6}\\yrs$) massive stars; \\item Tracing the physical conditions of (and processes occurring in) the environments where GRBs occur which hold clues for understanding the mechanism for making a burst, e.g., a flat or gray extinction law for GRB host galaxies would imply a dense circumburst environment where dust undergoes coagulational-growth or a preferential destruction of small grains; and \\item Probing the interstellar medium (ISM) of high-redshift galaxies and the cosmic star formation history -- because of their intense luminosity which allows their detection at cosmological distances, GRBs are a powerful tool to study the star formation history up to very high redshifts; e.g., the dust and extinction properties of GRB hosts would help understand the nature of dark bursts and the dark burst fraction which would place important constraints on the fraction of obscured star formation in the universe (e.g. see Djorgovski et al. 2001, Ramirez-Ruiz et al.\\ 2002). \\end{itemize} However, our current understanding of the dust extinction in GRB host galaxies is still very poor. Existing studies on this often draw conclusions in conflict with each other (see \\S\\ref{sec:status} for details). We argue that this could be caused by the prior adoption of a {\\it template} extinction law in fitting the observed GRB afterglow spectra to derive dust extinction (\\S\\ref{sec:status}). We propose in this work an alternative, robust method based on an analytical formula which can restore the widely adopted template extinction laws (\\S\\ref{sec:approach}). For illustration, we apply this approach to GRB\\,000301C and GRB\\,021004 (\\S\\ref{sec:test}). We demonstrate in \\S\\ref{sec:discussion} the uniqueness of the derived extinction laws. The robustness of this approach will be discussed in a separate paper (Liang \\& Li 2008a) in which the afterglow SEDs of $>$\\,50 GRBs of a wide range of properties are successfully modeled and for which the inferred extinction curves are diverse, with some differing substantially from any of the template extinction curves. ", "conclusions": "} We have also fitted the afterglow SEDs of GRB\\,000301C and GRB\\,021004 in terms of the MW, SMC, LMC, Calzetti, and ``linear'' % template extinction curves (see Table \\ref{tab:grbmod} and Figs.\\,\\ref{fig:GRB000301C},\\ref{fig:GRB021004}). Since for a given template extinction law the wavelength-dependence of the extinction $A_\\lambda/A_V$ is fixed, we are now left with only three parameters: $\\Fo$, $\\beta$, and $A_V$. The models based on the MW and LMC extinction laws could not fit the observed SEDs at all. This is because the 2175$\\Angstrom$ extinction feature which is prominent in the MW and LMC curves is absent in the SEDs of GRB\\,000301C and GRB\\,021004. In contrast, the SMC and ``linear'' models closely fit the afterglow SEDs of these two bursts, better than the Drude model proposed here as measured by $\\chi^2/N_{\\rm d.o.f.}$ (see Table \\ref{tab:grbmod}). While the Drude model has three more parameters than the SMC and ``linear'' models, the quality of the fitting of the Drude model is even not as good as that of the SMC or ``linear'' model. Then, why do not we simply adopt the SMC or ``linear'' model? First of all, we should note that there are no physical reasons for a prior assumption of a known extinction law, either that of the SMC, LMC, ``linear'' or MW: the composition and size distribution (and therefore the extinction law) of the dust in the dense circumburst clouds of GRB hosts with a wide range of metallicities and evolutionary stages are not expected to resemble that of the MW, LMC, or SMC (e.g. see Dwek 2005). In literature, a SMC-type extinction is often assumed for low-metallicity environments. However, there is no physical basis for this (except the lack of grain growth in these regions because of the lack of raw dust materials -- the SMC dust, on average, is substantially smaller than that of the Milky Way [see Weingartner \\& Draine 2001]). Moreover, it is known that the GRB hosts have a wide range of metallicities. Indeed, the reasons why the MW, LMC and SMC laws are often used for GRB afterglow SED modeling are mainly (1) little is known about the extinction laws of other galaxies, and (2) the Pei (1992) formula for the MW, LMC and SMC extinction laws is numerically convenient for computer implementation. Second, although the SMC-type extinction is preferred in most of the present afterglow SED modeling studies, only the Drude approach is capable of reproducing the SEDs of those reddened by gray extinction or by non-conventional extinction. Indeed, it was shown that the afterglow SED of GRB\\,050904 at a redshift of $z$\\,$\\approx$\\,6.3 cannot be explained by dust reddening with any of the conventional (MW, SMC, Calzetti) extinction curves; instead, it can be well reproduced by invoking the extinction curve inferred for a distant quasar at $z$\\,=\\,6.2 (Maiolino et al.\\ 2004), suggesting that the properties of dust may evolve beyond $z$\\,=\\,6 (Stratta et al.\\ 2007). Third, the Drude model would at least complement the models using template extinction curves, particularly for those bursts for which the Drude model gives a larger $\\chi^2/N_{\\rm d.o.f.}$ (but still fits the observed SEDs well). Given that the derived extinction $A_V$ and the intrinsic spectral slope $\\beta$ differ appreciably among different approaches (see Table \\ref{tab:grbmod}), the SMC model (and other models) should be used along side with the Drude model to gain insight into the ``true'' extinction and the ``true'' spectral slope. We finally demonstrate the uniqueness of the extinction curve inferred from the Drude approach. To this end, we generate three sets of afterglow ``photometry data'' by reddening the intrinsic afterglow spectrum $F_\\nu\\,(\\mu {\\rm Jy})$\\,=\\,$5.2\\times 10^8 \\left(\\nu/{\\rm Hz}\\right)^{-0.5}$ of a burst at $z\\approx 2$ respectively with three template extinction laws: MW, SMC, and Calzetti, each with $A_V=0.5\\magni$. We then apply the Drude approach to these three sets of artificially-created GRB afterglow data. As shown in Figure\\,\\ref{fig:test}, we uniquely restore the MW, SMC, and Calzetti extinction laws: the inferred extinction curves are almost identical to that used to redden the intrinsic spectrum (the derived parameters [see Table\\,\\ref{tab:test}] are essentially the same as those tabulated in Table\\,\\ref{tab:extcurv})." }, "0808/0808.3818_arXiv.txt": { "abstract": "We present new proper motions from the 10 m Keck telescopes for a puzzling population of massive, young stars located within 3\\farcs5 (0.14 pc) of the supermassive black hole at the Galactic Center. Our proper motion measurements have uncertainties of only 0.07 mas yr$^{-1}$ (3 km s$^{-1}$), which is $\\gtrsim$ 7 times better than previous proper motion measurements for these stars, and enables us to measure accelerations as low as 0.2 mas yr$^{-2}$ (7 km s$^{-1}$ yr$^{-1}$). Using these measurements, line-of-sight velocities from the literature, and 3D velocities for additional young stars in the central parsec, we constrain the true orbit of each individual star and directly test the hypothesis that the massive stars reside in two stellar disks as has been previously proposed. Analysis of the stellar orbits reveals only one of the previously proposed disks of young stars using a method that is capable of detecting disks containing at least 7 stars. The detected disk contains 50\\% of the young stars, is inclined by $\\sim115^\\circ$ from the plane of the sky, and is oriented at a position angle of $\\sim100^\\circ$ East of North. Additionally, the on-disk and off-disk populations have similar K-band luminosity functions and radial distributions that decrease at larger projected radii as $\\propto r^{-2}$. The disk has an out-of-the-disk velocity dispersion of 28 $\\pm$ 6 km s$^{-1}$, which corresponds to a half-opening angle of $7^\\circ \\pm 2^\\circ$, and several candidate disk members have eccentricities greater than 0.2. Our findings suggest that the young stars may have formed {\\it in situ} but in a more complex geometry than a simple, thin circular disk. ", "introduction": "\\label{sec:intro} The center of our Galaxy harbors not only a supermassive black hole \\citep[Sgr A*, $M_\\bullet \\sim 4 \\times 10^6$ \\msun;][]{ eckart96,genzel96,ghez98pm,ghez00nat,ghez03spec,ghez05orbits, schodel02,schodel03,eisenhauer06}, but also a population of massive (10-120 \\msun), young ($\\lesssim$10-100 Myr) stars whose existence is a puzzle. The origin of such young stars has been difficult to explain since the gas densities observed today are orders of magnitude too low for a gas clump to overcome the extreme tidal forces and collapse to form stars \\citep[e.g.][for reviews]{sanders92,morris93,ghez05orbits,alexander05review}. And yet, within the central parsec of our Galaxy, nearly 100 stars have been classified as OB main-sequence stars, more luminous OB giants and supergiants, and post-main-sequence Wolf-Rayet stars \\citep{allen90,krabbe91,blum95heI,krabbe95,tamblyn96, najarro97,ghez03spec,paumard06}, with the more evolved massive stars having ages as young as 6$\\pm$2 Myr \\citep{paumard06}. Populations of young stars have also been observed in the nuclei of other galaxies, such as M31 \\citep{bender05}, suggesting that star formation near a supermassive black hole may be a common, but not understood, phenomenon in galaxy evolution. The close proximity of the black hole at the center of the Milky Way provides a unique laboratory for studying this ''paradox of youth'' \\citep[e.g.][]{ghez03spec,ghez05orbits,schodel03,eisenhauer06}. Proposed resolutions to the paradox of youth can be grouped into several broad categories, including (1) rejuvenation of an older population such that older stars appear young, (2) dynamical migration from larger radii, and (3) {\\it in situ} formation. Rejuvenation scenarios include stripping \\citep{davies98,davies05} or tidal heating of the atmospheres of old stars \\citep{alexander03}, or combining multiple low mass stars via collisional mergers to form a higher-mass hot star akin to a ``blue straggler'' \\citep{lee96gcmergers,morris93,genzel03cusp}. Although these processes may be candidates for explaining the closest young stars within the central arcsecond, they cannot account for the OB giants, OB supergiants, and Wolf-Rayet stars that are located at larger radii (1\\arcsec-14\\arcsec), since the rate of collisions is too low to produce the observed total numbers. Thus, it appears that these massive young stars must have formed, or were deposited, in the central region within the last 4-8 Myr. Dynamical migration scenarios attempt to resolve the paradox of youth with the formation of a massive star cluster at larger distances from the black hole (3-30 pc). Such a cluster would spiral in due to dynamical friction and deposit stars at smaller radii where they are observed today \\citep{gerhard01}. However, for a cluster to reach the central parsec in only a few million years, it must be very massive and centrally concentrated \\citep{kim03,pzwart03irs16,mcmillan03,gurkan05}, and it may even require the existence of an intermediate-mass black hole (IMBH) as an anchor in the cluster core \\citep{hansen03,kim04}. {\\it In situ} star formation scenarios can resolve the paradox of youth if a massive, self-gravitating gas disk was once present around the black hole \\citep{levin03}. Such a disk would be sufficiently dense to overcome the strong tidal forces, and gravitational instabilities would then lead to fragmentation and the formation of stars, as has been suggested in the context of both the Galactic Center circumnuclear disk and AGN accretion disks in other galaxies \\citep[e.g.][]{kolykhalov80,shlosman89,morris96,sanders98,goodman03,nayakshinCuadra05}. Insight into the origins of the massive, young stars may be obtained through observations of the spatial distribution and stellar dynamics of this population. Already, high-resolution infrared imaging and spectroscopy have shown that the young stars between 0\\farcs5 and 14\\arcsec (0.02-0.6 pc) exhibit coherent rotation \\citep{genzel00}. Analyses of the statistical properties of the three-dimensional velocity vectors for these stars suggest that they may reside in two disks. The first proposed disk has a clockwise sense of rotation, as projected onto the plane of the sky \\citep[][hereafter: clockwise-rotating or CW disk]{levin03}, while the second proposed disk is counter-clockwise-rotating \\citep[CCW][]{genzel03cusp} and is nearly perpendicular to the first. The proposed disks extend from $\\sim$0\\farcs8 to at least 7\\arcsec \\citep{paumard06}. Other velocity vector analyses show that there are possible co-moving groups or clusters of stars, including the IRS 13 cluster, which is proposed to lie within the putative CCW disk \\citep{maillard04irs13,schodel05}, and the IRS 16SW co-moving group, which are also consistent with the proposed CW disk \\citep{lu05irs16sw}. The two proposed disks are inferred to be oriented with an inclination and angle to the ascending node of [$i_{CW}$=127$^\\circ \\pm$ 2$^\\circ$, $\\Omega_{CW}$=99$^\\circ \\pm$ 2$^\\circ$] and [$i_{CCW}$=24$^\\circ \\pm$ 4$^\\circ$, $\\Omega_{CCW}$=167$^\\circ \\pm$ 7$^\\circ$] and to have a finite angular thickness of $\\Delta\\theta_{CW} \\sim 14^\\circ$ and $\\Delta\\theta_{CW} \\sim 19^\\circ$ where $\\Delta\\theta$ is the standard deviation of the orbital inclinations distributed normally about the disk plane \\citep{paumard06}. The thickness of the stellar disks has been attributed to thickening as a result of gravitational interactions between the two disks, which provides an estimate of the disk masses \\citep{nayakshin06thick}. The derived mass is smaller than the mass inferred from the number of observed young stars, assuming a Salpeter initial mass function (IMF); accordingly, \\citet{nayakshin06thick} suggest that the disks have a top-heavy mass function. Both {\\it in situ} gas disk and in-spiraling star cluster formation scenarios have been used to explain the kinematics of this young star population and to predict that the stars should lie in a common orbital plane. However, the presence of two stellar disks with similarly aged populations requires either two nearly concurrent gas disks or two infalling star clusters; and both of these scenarios are difficult to produce. Therefore, to understand the recent star formation history, it is critical to measure the orbital planes of individual stars in order to confirm the existence of the two stellar disks previously derived from a statistical analysis of velocity vectors alone. The {\\it in situ} gas disk and inspiraling star cluster formation scenarios predict different structures and evolutions for the resulting stellar disk, particularly with respect to the eccentricities and radial distribution of stars within the disk. Early models of a self-gravitating gas disk around the supermassive black hole at the center of the Milky Way produce stars with a steep radial profile in the disk surface density, $\\Sigma \\propto r^{\\alpha}$, with $\\alpha \\sim -2$ \\citep{linPringle87,levin06}. These models typically result in stars on circular orbits as would be the case for the slow build up of a gas disk that is circularized before there is sufficient mass for gravitational instabilities to set in \\citep{milosav04,nayakshinCuadra05,levin06}. The stellar eccentricities of an initially circular disk can relax to higher eccentricities up to $e_{rms} = \\sqrt{} \\sim $0.15 for a normal IMF or $e_{rms} \\sim $0.3 for a top-heavy IMF \\citep{alexander07imf,cuadra08}. More recent models have also shown that star formation can occur rapidly before circularization in an initially eccentric disk as might result from the infall of a single massive molecular cloud or a cloud-cloud collision \\citep{sanders98,nayakshin07sims,alexander08}. These eccentric self-gravitating accretion disk models typically produce a more top-heavy IMF than initially circular disks. On the other hand, an inspiraling star cluster would dissolve into a disk of stars with a flatter radial profile \\citep[$\\Sigma \\propto r^{-0.75}$;][]{berukoff06} whose orbital eccentricities would reflect the eccentricity of the cluster's orbit, which could be either circular or eccentric \\citep{pzwart03irs16,mcmillan03,kim03,kim04,gurkan05,berukoff06}. Previous measurements of the radial distribution of young stars yields a steep radial profile consistent with {\\it in situ} formation \\citep{paumard06}. Also, the eccentricities of the stars have previously been estimated from observations by assuming that the stars orbit in a disk; however, there are conflicting results claiming that the stars in the clockwise-rotating disk are on nearly circular orbits \\citep{paumard06} or on eccentric orbits \\citep{beloborodov06}. Determining the radial profile and stellar eccentricities of stars in a disk may provide observational constraints on the origin of the young stars. We present an improved proper motion study that yields an order of magnitude more precise proper motions and the first measurement of accelerations in the plane of the sky for stars outside the central arcsecond. By combining the stellar positions, proper motions, radial velocities, and accelerations, we estimate stellar orbital parameters and test whether the young stars reside on one or two stellar disks in a more direct manner than previous methods using only velocity information. This provides a {\\it direct} test of the existence, membership, and properties of these disks. The observations are described in \\S\\ref{sec:obs} and the astrometric analysis procedure and results are detailed in \\S\\ref{sec:astrometry}. Orbit analysis and results are presented in \\S\\ref{sec:orbitAnalysis} and \\S\\ref{sec:orbitResults} and a discussion of the implications for the origin of the massive, young stars at the Galactic Center is presented in \\S\\ref{sec:discussion}. ", "conclusions": "In summary, the advent of laser guide star adaptive optics has allowed us to retroactively improve our 11 year astrometric data set used for monitoring stars orbiting our Galactic Center. This has increased our proper motion precision, with resulting uncertainties of $\\sim$3 km s$^{-1}$, and allowed us, for the first time, to make measurements of and place limits on accelerations for stars outside the central arcsecond out to a radius of 3\\farcs5, with typical 3$\\sigma$ acceleration limits of -0.19 mas yr$^{-2}$. By combining our improved stellar positions and proper motions with radial velocity information from the literature, we compute orbits for individual young stars proposed to lie in stellar disks orbiting the supermassive black hole. The orbits for the young stars confirm only a single disk of young stars at a high inclination rotating in a clockwise sense and there is no statistically significant evidence for a second disk. Stars within the well-defined, clockwise disk have an out-of-the-disk velocity dispersion of 28 $\\pm$ 6 km s$^{-1}$ and several stars have high eccentricities. These disk properties suggest that star formation may have occurred in a single event, rather than the two events previously needed to explain two stellar disks; however, there are open questions as to how $\\sim$50\\% of all young stars can be perturbed out of the disk plane and whether the apparent compact cluster, IRS 13, which is not part of the stellar disk, requires a separate star formation or dynamical event. Future directions include (1) obtaining new LGSAO data sets with improved astrometry to measure accelerations for the young stars at all radii and (2) identifying new young stars within the central parsec in order to better constrain the orbital properties of these stars and to study in detail the distribution of eccentricities and semi-major axes for stars both in and out of the disk." }, "0808/0808.2733_arXiv.txt": { "abstract": "{A large number of AGN have been monitored for nearly 30 years at 22, 37 and 87 GHz in Mets\\\"ahovi Radio Observatory. These data were combined with lower frequency 4.8, 8.0 and 14.5 GHz data from the University of Michigan Radio Astronomy Observatory, higher frequency data at 90 and 230 GHz from SEST, and supplementary higher frequency data from the literature to study the long-term variability of a large sample of AGN. Both the characteristics of individual flares from visual inspection and statistically-determined variability timescales as a function of frequency and optical class type were determined. Based on past behaviour, predictions of sources expected to exhibit large flares in 2008--2009 appropriate for study by GLAST and other instruments are made. The need for long-term data for properly understanding source behaviour is emphasised.} \\FullConference{Workshop on Blazar Variability across the Electromagnetic Spectrum\\\\ April 22-25, 2008\\\\ Palaiseau, France} \\begin{document} ", "introduction": "Active galactic nuclei (AGN) are variable across the whole electromagnetic spectrum. The radio regime is of special interest because we can directly observe the synchrotron radiation from the jet. The total flux density variations and flares seen in the flux curves are usually explained with shock-in-jet models where a disturbance creates a shock moving down in the jet, and as the shock develops we see a flare evolving from higher submm- and mm wavelengths towards lower radio frequencies \\cite{marscher85, hughes85}. We have used a sample of 90 AGN to study their long-term variability timescales \\cite{hovatta07, hovatta08b} and flare characteristics \\cite{hovatta08a, nieppola08}. Our extensive database enables us to study the correspondence between the shock model and the observations, and also the statistical differences between the different AGN types (27 high polarisation quasars (HPQs), 33 low polarisation quasars (LPQs), 25 BL Lacertae objects (BLOs), and 5 radio galaxies (GALs)). Our sample includes bright sources that have flux density at least 1 Jy in the active state. ", "conclusions": "We studied the long-term radio variability of a sample of 90 sources using statistical timescale analysis methods and visual inspection of the flare parameters. Our main results are the following: \\begin{itemize} \\item{Fourier-based methods the DCF and the periodogram give very similar results as wavelets, but wavelets should be used when quasi-periodities are studied because they give information on the locality of the timescale. With wavelets it is possible to see if a timescale is long-lasting or just a short transient phenomenon in the flux curve. DCF or periodograms can then be used to verify the timescale more accurately.} \\item{Variability behaviour is complex and no clear periodicities could be found at radio frequencies. Episodes of quasi-periodic behaviour are common, and therefore false periodicities may be found if the temporal coverage is inadequate.} \\item{Flares are seen, on average, every 4 years in all the source types at 37\\,GHz but when intrinsic redshift-corrected timescales are studied, the quasars have shorter timescales of 2 years compared to the 3-4 years of BLOs. This could indicate that shocks are produced less frequently in BLOs than in quasars.} \\item{Median duration of a flare is 2.5 years at 22 and 37\\,GHz, but the range in durations is between 0.3 and 13.2 years. When comparing the duration with intrinsic redshift- and Doppler-corrected peak luminosities, we found that the energy release in a flare does not increase with the duration of the flare.} \\item{Flares adhere quite well to the predictions of the shock model but the scatter in the data, due to poor sampling and complicated structure of the flares, is still large.} \\item{By combining the median duration of flares, 2.5 years, with the average time between the flares, 4 years, we see that multifrequency campaigns should last for 5-7 years in order to catch the source in both its highest and lowest activity states.} \\item{Long-term monitoring is essential in understanding the true behaviour of these sources at radio frequencies.} \\end{itemize}" }, "0808/0808.2505_arXiv.txt": { "abstract": "We present a comprehensive abundance analysis of 27 heavy elements in bright giant stars of the globular clusters M4 and M5 based on high resolution, high signal-to-noise ratio spectra obtained with the Magellan Clay Telescope. We confirm and expand upon previous results for these clusters by showing that (1) all elements heavier than, and including, Si have constant abundances within each cluster, (2) the elements from Ca to Ni have indistinguishable compositions in M4 and M5, (3) Si, Cu, Zn, and all $s$-process elements are approximately 0.3 dex overabundant in M4 relative to M5, and (4) the $r$-process elements Sm, Eu, Gd, and Th are slightly overabundant in M5 relative to M4. The cluster-to-cluster abundance differences for Cu and Zn are intriguing, especially in light of their uncertain nucleosynthetic origins. We confirm that stars other than Type Ia supernovae must produce significant amounts of Cu and Zn at or below the clusters' metallicities. If intermediate-mass AGB stars or massive stars are responsible for the Cu and Zn enhancements in M4, the similar [Rb/Zr] ratios and (preliminary) Mg isotope ratios in both clusters may be problematic for either scenario. For the elements from Ba to Hf, we assume that the $s$- and $r$-process contributions are scaled versions of the solar $s$- and $r$-process abundances. We quantify the relative fractions of $s$- and $r$-process material for each cluster and show that they provide an excellent fit to the observed abundances. ", "introduction": "\\label{sec:intro} Globular clusters continue to play a vital role in testing many aspects of stellar evolution and stellar nucleosynthesis. Observational and theoretical studies of globular clusters have focused heavily upon (1) the star-to-star light element abundance variations \\citep{cottrell81,langer95,gratton04}, (2) the cluster-to-cluster variation in the color distribution of horizontal branch stars, the so-called ``2nd parameter effect'' \\citep{sandage67,lee94,carretta06}, and (3) the multiple populations as inferred from large spreads in metallicity and/or detailed structure in color-magnitude diagrams \\citep{butler78,norris95b,bekki06}. Abundance measurements of the $s$-process and $r$-process elements offer great insight into stellar nucleosynthesis and globular cluster chemical evolution. Aside from M15, a metal-poor cluster that displays a scaled-solar $r$-process abundance distribution \\citep{sneden97,sneden00b,otsuki06}, in general only a handful of $s$-process elements (e.g., Y, Zr, Ba, La) and the $r$-process element Eu have been measured in globular clusters. The globular clusters M4 and M5 are particularly well suited for refining our understanding of stellar evolution and stellar nucleosynthesis, especially for the neutron-capture elements. \\citet{M4,M5} showed that these clusters have essentially identical metallicities, [Fe/H] = $-$1.2, based on high resolution spectra of large samples, 36 stars in each cluster. They also showed that the abundance similarities for these clusters extend to numerous $\\alpha$- and Fe-peak elements as well as the $r$-process element Eu. However, the $s$-process elements revealed striking abundance differences between these two clusters. Specifically, the heavy $s$-process elements Ba and La are overabundant in M4 relative to M5 \\citep{M4,M5}. With the notable exception of $\\omega$ Cen, M4 may be uniquely enriched in $s$-process elements among the Galactic globular clusters \\citep{pritzl05}. \\citet{rbpbm4m5} recently extended the analysis of neutron-capture elements in these two clusters to Rb and Pb, $s$-process elements which may be overproduced in metal-poor asymptotic giant branch (AGB) stars \\citep{busso99,travaglio01}. Whereas Pb production is dominated by 2 to 4$M_{\\odot}$ low-metallicity AGB stars \\citep{travaglio01}, the intermediate-mass 4 to 8$M_{\\odot}$ AGB stars are predicted to dominate Rb production \\citep{vanraai08}. Not surprisingly, M4 again had higher abundance ratios [Rb/Fe] and [Pb/Fe] than M5. However, the abundance ratios [Rb/X] for X = Y, Zr, and La were very similar in the two clusters indicating that the nature of the $s$-process products is very similar for both clusters but that M4 formed from gas with a higher concentration of these products. A comprehensive study of $s$-process elements in these two clusters promises to provide a novel observational study of the $s$-process at low metallicities as well as valuable clues to the chemical evolution diversity of globular clusters. In this paper we present such an analysis, focusing upon a suite of $\\alpha$-, Fe-peak, $s$-process, and $r$-process elements. ", "conclusions": "\\label{sec:summary} In this paper we present abundance ratios [X/Fe] for a large number of $\\alpha$-, Fe-peak, $s$-process, and $r$-process elements for 12 bright giants in the globular cluster M4 and 2 bright giants of the globular cluster M5. This comprehensive abundance analysis is only possible due to the large wavelength coverage, high resolution, and very high S/N spectra. For all elements in this study, we find no evidence for star-to-star abundance variations in either cluster. We confirm and extend upon previous results for these clusters by showing that (1) for the elements from Ca to Ni, M4 and M5 have identical abundance ratios, (2) M4 shows overabundances by roughly 0.3 dex for Si, Cu, Zn, and all $s$-process elements relative to M5, and (3) for the $r$-process elements, M5 may have slightly higher abundances than M4 by 0.1 dex. We also measure Mg isotope ratios and find that the ratios are solar in both clusters, with no sign of any star-to-star variation within each cluster. The ratios $^{25}$Mg/$^{24}$Mg and $^{26}$Mg/$^{24}$Mg exceed values found in field halo stars at the same metallicity, e.g., Gmb 1830, which implies differences in the clouds from which globular clusters and field halo stars formed. However, we regard these ratios as preliminary since the spectral resolution was insufficient to accurately distinguish $^{25}$Mg from $^{26}$Mg. There is no clear explanation for the M4-M5 Si abundance differences since the abundances of Ca should, but do not, follow the behavior of Si. For the elements from Ba to Hf, we find that the mean abundances in M4 and M5 are well explained by scaled versions of the solar $s$- and $r$-process abundances, albeit with different mixes of $s$- and $r$-process material for each cluster. Therefore, no new $s$-process site is required to explain the M4-M5 abundance differences for the elements from Ba to Hf. However, although the Th abundances lie above these predictions, the ratio [Th/Eu] is identical in both clusters indicating that the universality of the $r$-process extends to Th in these clusters and that no differential decay of Th has occurred, i.e., the clusters have identical ages. The Pb abundances lie below the predictions, by different amounts for each cluster. Therefore, the sources of the $s$-process may differ between M4 and M5, at least regarding the production of Pb via the strong component. The abundance differences between M4 and M5 for Cu and Zn are particularly intriguing given that their nucleosynthetic origins continue to be debated. The $s$-process elements, produced in AGB stars (and massive stars), share a similar abundance behavior to Cu and Zn in M4 and M5. Updated, but preliminary, yields from AGB models indicate that small amounts of Zn may be produced only in the most massive AGB stars. These models also predict that the most massive AGB stars may produce Cu in contrast to our current understanding of Cu production. Massive AGB stars are expected to produce large amounts of the neutron-rich Mg isotopes, and observations of high $^{26}$Mg/$^{24}$Mg ratios in field and cluster stars confirm the AGB yields. However, preliminary measurements show no difference in the Mg isotope ratios between these two clusters which constrains the contribution of intermediate-mass AGB stars to the Cu and Zn enhancements in M4. Si, which is produced in massive stars, shows a similar abundance behavior to Cu and Zn in M4 and M5. Cu and Zn may be produced in massive stars via the weak $s$-process. While the abundance ratios [Rb/Sr], [Rb/Y], and [Rb/Zr] are predicted to increase via the weak $s$-process, our measurements do not reveal any cluster-to-cluster variations in these abundance ratios which suggests that either massive stars are not responsible for the Cu and Zn differences or that metal-poor massive stars do not alter the [Rb/Zr] ratio. Of great interest would be detailed chemical evolution modeling of these two clusters to gain insight into the origin of the Cu and Zn abundance differences and therefore their nucleosynthesis production sites." }, "0808/0808.3579_arXiv.txt": { "abstract": "{ Hydromagnetic stresses in accretion discs have been the subject of intense theoretical research over the past one and a half decades. Most of the disc simulations have assumed a small initial magnetic field and studied the turbulence that arises from the magnetorotational instability. However, gaseous discs in galactic nuclei and in some binary systems are likely to have significant initial magnetisation. Motivated by this, we performed ideal magnetohydrodynamic simulations of strongly magnetised, vertically stratified discs in a Keplerian potential. Our initial equilibrium configuration, which has an azimuthal magnetic field in equipartion with thermal pressure, is unstable to the Parker instability. This leads to the expelling of magnetic field arcs, anchored in the midplane of the disc, to around five scale heights from the midplane. Transition to turbulence happens primarily through magnetorotational instability in the resulting vertical fields, although magnetorotational shear instability in the unperturbed azimuthal field plays a significant role as well, especially in the midplane where buoyancy is weak. High magnetic and hydrodynamical stresses arise, yielding an effective $\\alpha$-value of around 0.1 in our highest resolution run. Azimuthal magnetic field expelled by magnetic buoyancy from the disc is continuously replenished by the stretching of a radial field created as gas parcels slide in the linear gravity field along inclined magnetic field lines. This dynamo process, where the bending of field lines by the Parker instability leads to re-creation of the azimuthal field, implies that highly magnetised discs are astrophysically viable and that they have high accretion rates. ", "introduction": "Since the seminal work of \\cite{BalbusHawley1991} it has been widely recognised that hydromagnetic stresses play a central role in the dynamics of Keplerian gaseous discs. Much of the subsequent theoretical work has been devoted to the study of the magnetic dynamo driven by a combination of magnetorotational instability (MRI), rotation, and Keplerian shear \\citep[][see \\citealp{BrandenburgSubramanian2005}, for an excellent review of astrophysical dynamo theory]{Brandenburg+etal1995,Hawley+etal1996}. Typically, numerical simulations of the MRI-driven dynamo begin with an initial zero net flux magnetic field with an associated pressure which is a small fraction of the thermal pressure. The idea is that the MRI-driven turbulence should increase the characteristic coherence length of the magnetic field, and grow its mean strength to a significant fraction of the equipartition value. The simulations, however, have had at best a mixed success in explaining high and persistent inflow rates observed in astrophysical accretion discs. Firstly, a number of recent numerical experiments have shown that the effectiveness of the dynamo depends in a critical way on the value magnetic Prandtl number, defined as the ratio of the collisional kinematic viscosity to the magnetic diffusivity \\citep[][following suggestive non-disc simulations of \\citealp{Schekochihin+etal2005} and a careful study of previously published MRI-turbulence results by \\citealp{Pessah+etal2007}]{LesurLongaretti2007,Fromang+etal2007}. The upshot of this work is that when the Prandtl number is significantly less than one, as is expected in most accretion discs \\citep{BalbusHenri2008}, the dynamo seems to fail\\footnote{Also the magnitude of the turbulent magnetic stresses decrease as the grid resolution increases (decreasing the numerical diffusivity accordingly), with no convergence in sight \\citep{FromangPapaloizou2007}. In may well be, however, that this problem is related to an insufficient size of the simulation domain, and that bigger, stratified shearing boxes will show better convergence \\citep{RegevUmurhan2008}.}. Secondly, in the zero net flux simulations where all parameters are chosen so that the dynamo works, the measured value of the Shakura-Sunyaev parameter $\\alpha$ is typically of order $10^{-3}$ and at most $10^{-2}$ \\citep[e.g.][]{Brandenburg+etal1995,Sano+etal2004,JohansenKlahr2005,FromangNelson2006}. This is $1$--$2$ orders of magnitude smaller than what is required to explain the high accretion rates in dwarf novae systems \\citep{King+etal2007}. The weak initial magnetic fields assumed in almost all disc-MRI simulations may be unrealistically small for many astrophysical discs. We give here two examples:\\newline \\hskip .1in 1. Extended accretion discs around central supermassive black holes \\citep[such as the ones traced by megamasers in nearby galaxies; see e.g.][]{Greenhill2007,Vlemmings+etal2007} are fed by molecular material in the interstellar medium. Molecular clouds are known to have large scale superthermal magnetic fields, and thus the initial magnetic fields in AGN discs are likely to be comparable to or larger than the thermal equipartition values. The $\\sim$2 pc molecular circumnuclear disc in our galactic centre is permeated by large scale equipartition magnetic fields \\citep{WardleKonigl1990,Hildebrand+etal1990} which are certain to play an important dynamical role in its subsequent evolution.\\newline \\hskip .1in 2. Accretion discs in close binary systems, which are fed by a Roche-lobe overflow from a tidally distorted low mass star, may be initially magnetised. The magnetisation $1/\\beta$ of the gas as it departs from the donor star can be estimated as \\begin{equation}\\label{binary1} {1\\over \\beta}\\sim {B_\\star^2 \\over 8\\pi \\rho c_s^2}, \\end{equation} where $B_\\star$ (measured in Gauss) is the stellar magnetic field at the surface, $c_s$ is the speed of sound at the stellar surface, and $\\rho$ is the density of gas as it becomes detached from the star. The density $\\rho$ can be estimated as \\begin{equation}\\label{binary2} \\rho\\sim {\\dot{M}\\over 4\\pi R_\\star^2 c_s}, \\end{equation} where $\\dot{M}$ is the average accretion rate of the companion, and $R_\\star$ is the radius of the star. We get \\begin{equation} {1\\over \\beta}\\sim 0.3 \\left({\\dot{M}\\over 10^{-9}M_{\\odot}/\\hbox{yr}}\\right)^{-1} \\left({c_s\\over 3\\hbox{km/s}}\\right)^{-1}\\left({B_\\star \\over 1\\hbox{G}}\\right)^2 \\left({R_\\star\\over 10^{11}\\hbox{cm}}\\right)^2. \\label{binary3} \\end{equation} The streaming and shearing motion after the gas detaches from the star can further amplify the magnetic field. Clearly, there is a realistic parameter range where the initial magnetisation is high. Motivated by these astrophysical considerations, in this paper we perform numerical experiments on gas discs which contain initially strong magnetic fields, with magnetic pressure comparable to that of the gas. Specifically, we initialise the azimuthal field so that it is subject to the Parker instability \\citep[PI,][]{Parker1966} in the vertical stratification. In terms of initial conditions our physical set up is close to \\cite*{Machida+etal2000}. However, \\cite{Machida+etal2000} focused on the formation and evolution of a disc corona, and the spirit or their numerical experiment was different from ours. They have simulated the whole circular disc, and have introduced reflecting boundary conditions at the midplane of the disc, which suppresses field anchoring in the midplane. By contrast, our purpose is to understand the long-term dynamics of the disc, possible field confinement, and the dynamo processes related to strong magnetic fields. Therefore, our shearing-box simulations focus on a small part of the disc and study it with high numerical resolution. We simulate the fluid both below and above the midplane, and thus have no artificial boundary conditions at the midplane of the disc. We find that both the short term and, more importantly, the long term behaviour of initially strongly magnetised discs is radically different from that of their weakly magnetised counterparts. We observe the following three-step dynamics: (a) the Parker instability expels azimuthal field in huge arcs, creating vertical field which becomes the seed for a strong magnetorotational instability, (b) matter sliding down inclined field lines stretches the azimuthal magnetic field and creates a vertically dependent large scale mean radial field, and (c) the Keplerian shear recreates azimuthal field from the stretching of the radial field. The latter step closes the dynamo cycle, in much the same way as was sketched by \\cite{ToutPringle1992} a decade and a half ago, although non-axisymmetric magnetorotational instability in the azimuthal field also plays an important role in creating accretion stresses in our simulations \\citep{BalbusHawley1992,FoglizzoTagger1995,TerquemPapaloizou1996}. The azimuthal field remains strongly concentrated towards the disc midplane; this is in contrast with the simulations of Parker instability which do not include strong Keplerian shear in the radial direction \\citep{Kim+etal1998}. We show that the dynamo is robust and stable over at least tens of orbital periods, and that the accretion torque {\\it increases} if we use a finer grid. We observe $\\alpha$-values as high as $0.1$ in our highest resolution run. The plan of the paper is as follows. In \\S\\ref{s:numerics} we detail the mechanics of our numerical experiments, and in the following section (\\S\\ref{s:noshear}) we test our computer code by comparing the results of Parker instability simulations without Keplerian shear to the extensive literature that exists on the PI under rigid rotation. In \\S\\ref{s:shear} we perform simulations with the Keplerian shear included, and present our main results. We devote the next section (\\S\\ref{s:confinement}) to analysing the confinement of azimuthal flux to the disc by a dynamo process. In \\S\\ref{s:conclusions} we conclude with the discussion of the astrophysical implications and possible future improvements of our work. ", "conclusions": "\\label{s:conclusions} In this paper we consider the evolution of strongly magnetised Keplerian accretion discs. Our numerical experiments show that the hydromagnetic state of the gas flow is very different from what is seen in zero net flux simulations. The Parker instability leads to huge magnetic arcs rising several scale heights from the disc midplane, and the magnetorotational instability in turn feeds off the vertical fields and creates a highly turbulent flow, an interaction that was predicted analytically by \\cite{ToutPringle1992}. Although the flow is stochastic and time fluctuating, we have identified an underlying dynamo process that couples the vertically dependent mean radial and azimuthal magnetic field components. As gas slides down inclined field lines, it obtains a helical motion due to Coriolis forces, and thus the azimuthal field lines are twisted in such a way as to create a mean electromotoric force in the direction of the unperturbed field line -- a configuration prone to create radial field. In turn the large scale radial field regenerates the azimuthal field through Keplerian shear. Although Parker instability dominates the linear growth phase, we have found evidence for magnerotational instability in the azimuthal field as well. In the midplane of the disc, where the buoyancy is weak, azimuthal MRI drives the initial evolution towards turbulence \\citep{FoglizzoTagger1994,FoglizzoTagger1995,TerquemPapaloizou1996}. These two related instabilities, magnetorotational instability in the vertical fields created by the Parker instability and magnetorotational swing instability in the azimuthal fields, both rely on azimuthal flux confinement and can coexist in the linear as well as in the non-linear state of transmagnetic accretion discs. Such a path to accretion, based on the interaction of Parker and magnetorotational instabilities, has at least two appealing traits. First of all that the vertical fields that feed the magnetorotational instability are created in a transparent way by the Parker instability. Zero magnetic flux models must most likely rely on a small scale dynamo in order to create vertical fields, and there is mounting evidence that such a dynamo would not operate in the bulk part of accretion discs where the magnetic Prandtl number is much lower than unity \\citep{Schekochihin+etal2005,Fromang+etal2007}. The second appealing result of our model is that the Maxwell and Reynolds stresses are significant ($\\alpha\\approx0.1$). Such high accretion stresses could solve the problem that observed accretion rates are often one or two orders of magnitude higher than the accretion rates obtained in zero net flux MRI simulations \\citep{King+etal2007}. The regeneration of azimuthal field by the shearing of an appropriate radial field was seen in all our simulations that included Keplerian shear. As magnetohydrostatic equilibrium is compromised by the Parker instability, gas streams down along inclined field lines. Coriolis force diverts the gas to the right, and a radial magnetic field is created as the azimuthal field is subjected to shear-regions typically the size of the Parker instability. Eventually magnetic reconnection leads to a coherent large scale radial magnetic field. This dynamo was predicted by \\cite{Parker1992} and subsequently observed in the rigid rotation simulations of \\cite{Hanasz+etal2002}. To our knowledge we are the first to point out the relevance of Parker's fast galactic dynamo to accretion discs and how it closes the accretion loop by replenishing the azimuthal field that is lost by magnetic buoyancy. The fine-tuned initial conditions with a purely azimuthal magnetic field and a constant ratio of magnetic to thermal pressure may be questioned. However our experiments with a combined azimuthal and vertical field shows that the Parker instability is robust even if the azimuthal field coexists with a moderately strong net vertical field, and that the additional vertical field component may indeed increase the accretion rate further. Our results may also be relevant for star formation in the galactic centre. Although there is currently no coherent accretion disc structure, the population of young, massive stars in a disc-like structure close to the galactic centre points towards the brief existence of an accretion disc some million years ago \\citep{LevinBeloborodov2003,MilosavljevicLoeb2004,Nayakshin+etal2007,Alexander+etal2008}. The disc was likely to be initially strongly magnetised, as indicated by the current high magnetisation of the circumnuclear molecular ring. Hence the type of MHD processes studied in this paper may be of central significance for the disc dynamics. The presented simulations of transmagnetic ($\\beta\\sim1$) discs argues that discs dominated by magnetic pressure $\\beta \\ll 1$ are astrophysically viable. The existence of such discs was conjectured by \\cite{Pariev+etal2003} and their limitations and observational consequences were explored by \\cite{BegelmanPringle2007}. A number of problems in accretion disc theory are alleviated by the presence of super-equipartition magnetic fields, among them is the long-standing issue of self-gravity and fragmentation of AGN discs \\citep{Goodman2003}. Future research into strongly magnetised accretion discs should also focus on the self-consistent modelling of the magnetisation of the material that feeds accretion discs in various environments and on the evolution of magnetised disc coronae, in light of effects such as reconnection, shearing of foot points, Ohmic heating and radiative cooling that take place there." }, "0808/0808.1056_arXiv.txt": { "abstract": "A rudimentary calculation is employed to evaluate the possible effects of $\\beta$-decays of excited state nuclei on the astrophysical r-process. Single particle levels calculated with the FRDM are adapted to the calculation of $\\beta$-decay rates of these excited state nuclei. Quantum numbers are determined based on proximity to Nilson model levels. The resulting rates are used in an r-process network calculation in which a supernova hot-bubble model is coupled to an extensive network calculation including all nuclei between the valley of stability and the neutron drip line and with masses 1$\\le$A$\\le$283. $\\beta$-decay rates are included as functional forms of the environmental temperature. While the decay rate model used is simple and phenomenological, it is consistent across all 3700 nuclei involved in the r-process network calculation. This represents an approximate first estimate to gauge the possible effects of excited-state $\\beta$-decays on r-process freezeout abundances. ", "introduction": "\\label{Introduction} The r-process is responsible for the synthesis of roughly half of all nuclei heavier than A$\\sim$70 and all of the actinides \\cite{cowan90,wallerstein97}. The solar system r-process abundances act as the canonical constraint to r-process theories as well as the prime indicator of the success of r-process models. Several r-process sites have been proposed; the hot-bubble region of a type II supernova (SNII) has been modeled fairly successfully. The composition of the environment in which the r-process occurs might be expected to have a profound effect on the final abundance distribution. Observations indicate that the r-process site is primary \\cite{sneden98} and further evidence may suggest that the r-process is also unique \\cite{cowan99}; it may occur in a single site or event. The uniqueness of the r-process site, however, remains a subject of study \\cite{qian98}. Nuclear properties also constrain the r-process, and the purpose of this work is the examination of one particular characteristic - $\\beta$-decay - as it relates to the r-process. The $\\beta$-decay inputs, and other nuclear physics inputs, have been shown \\cite{woosley94,meyer95} to have important effects on the success or failure of r-process models. This is somewhat unfortunate, as properties of only a few nuclei on the neutron closed shells closer to stability have been experimentally determined, while data for the rest are relegated to calculation \\cite{cowan90}. Of paramount importance is the determination of nuclear masses and $\\beta$-decay rates. Nuclear mass formulae based on the microscopic properties of nuclei are slowly replacing the empirical droplet models, and these can change resulting reaction rates by factors as large as 10$^8$ \\cite{goriely96}. As well, the r-process path is affected by the choice of mass formula, since the path roughly follows a line of constant S$_n$ \\cite{cowan90}. A successful r-process calculation must predict an abundance peak for nuclei in the A$\\sim$195 region - a difficulty over much of the r-process parameter space. However, as discussed below, $\\beta$-decays from excited state nuclei may help to mitigate this difficulty. They do so by allowing the r-process to proceed at a faster rate, thereby enhancing the abundances at higher masses. For the purposes of this study, the most recent semi-gross theory of $\\beta$-decay \\cite{nakata97} has been adapted to neutron-rich nuclei relevant to the r-process. The ability of this model to determine decay properties of an extremely wide range of nuclei with reasonable accuracy and speed makes it ideal for this preliminary calculation. In particular, the semi-gross theory has good agreement for very neutron-rich nuclei \\cite{ichikawa05,asai99}. It has also been used to improve the accuracy of decay rates for astrophysical calculations by incorporating first-forbidden transition strengths \\cite{moller03}. In its original form, the gross theory of $\\beta$-decay assumed that the energy states of a nucleus consist of a smoothed distribution with transition strengths that peak at or near the energy of the isobaric analog state \\cite{takahashi72,takahashi69}. Subsequent evolutions of the gross theory incorporated strength functions allowing for transitions of higher forbiddenness \\cite {takahashi71}, as well as improvements over the original theory to include odd-odd effects \\cite{nakata95}, sum rules \\cite{tachibana90}, even-odd mass differences \\cite{koyama70}, and improvements on the strength functions \\cite {kondoh85}. While the precision of the Fermi gas model used is consistent with that of the gross theory itself, more accurate models might be used. Recently, shell effects in the parent nucleons have been taken into account by using an energy distribution for single-particle states \\cite{nakata97,kondoh76}; this is denoted as the ``semi-gross theory.'' In these models, the energy distribution of the daughter states is still assumed to be smooth. Thus, the transition strength functions depend on the quantum numbers only of the initial parent states and are independent of the energy of the parent state; transition types are based on a statistical weight for a particular parent state to make a particular type of transition. The advantage of using single-particle strength functions is that quantum numbers can be assigned to the states easily, lending a better notion of the actual strength of each type of transition involved. Since decay rates of nearly all of the nuclei along the r-process path have yet to be studied in a laboratory, r-process calculations rely heavily on calculated decay rates. Further, the temperature of the r-process environment (~10$^9$K) necessitates accounting for nuclei in excited states, especially given the expected high level density of these far-from-stability nuclei. Some of the effects that might be expected if one considers excited state nuclei in r-process simulations include increased (n,$\\gamma$) and ($\\gamma$,n) rates, which might shift the r-process path, but would tend to counteract each other as the r-process is generally presumed to proceed at (n,$\\gamma$)$\\leftrightarrow$($\\gamma$,n) equilibrium, increased neutrino spallation rates, tending to enhance smoothing in post-processing, and increased beta-decay rates. However, at signicant excitation, neutron separation energies are low enough that neutron emission may be a dominant decay mode. This is discussed briefly along with the discussion of excited-state decays in the network calculation. Section \\ref{grosstheory} of this paper is an overview of $\\beta$-decay calculations and average properties in calculating decay rates calculated using the more recent semi-gross theory. This includes a brief review of the semi-gross theory of $\\beta$-decay and a description of the calculations of single-particle states and their relationship to $\\beta$-decay in \\S\\ref{spstates}. Energy levels are calculated in \\S\\ref{spcalc} using the Finite Range Droplet Model (FRDM). Results of $\\beta$-decay calculations of excitated state nuclei with the adapted FRDM are discussed in \\S\\ref{results}, along with their potential astrophysical importance. Application of these results to an r-process network calculation is made in \\S\\ref{r-process} with results from a model with several environmental parameter sets described in \\S\\ref{network_results}. Future work and possibilities are discussed in \\S\\ref{conclusion}, along with experimental possibilities. ", "conclusions": "\\label{conclusion} This work provides a study of the effects of excited state $\\beta$-decays on the r-process. A preliminary method was used to evaluate the possible effects of $\\beta$-decay rates of excited-state nuclei. Though the accuracy of the model is limited by the knowledge of single particle levels, an approximate treatment allows one to gauge the magnitude of effects on the r-process and provide impetus for further study. An empirical calculation was employed to find single-particle levels, and quantum numbers were deduced based on the level proximity to those of the spherical shell model. Although minor effects were found to result from inclusion of the excited state decays, there are also other possible effects that the inclusion of excited state nuclei may have on the r-process. One can imagine that if the excitation is due to the promotion of neutrons to higher-lying single-particle orbitals, then the photoneutron Q-value will decrease, and the ($\\gamma$,n) reaction rate may increase. Thus the effect of an increased rate might be to shift the r-process path closer to $\\beta$-stability. All models used in this calculation do a reasonable qualitative job of reproducing the solar abundance distribution for the mass region $80\\le{A}\\le{130}$. The fact that the abundance at low mass is roughly independent of the type of model used (hot or cold) is an indicator that the nuclei in this region have decay rates not as heavily dependent on temperature as some of the higher mass nuclei. In the more massive nuclei, the sensitivity of the decay rates on temperature resulted in a more pronounced shift in the path as the r-process evolves through freezeout. As mentioned previously, the dynamical treatment of the r-process is an important factor here in that the path continues to evolve even during freezeout, a result of $\\beta$-delayed neutron emission. With the mass formula used in this evaluation, it was found that the low mass nuclei have a higher probability of emitting two neutrons during $\\beta$-decay than the higher mass nuclei, which have a higher probability of emitting a single neutron during $\\beta$-decay. These available neutrons are recaptured, with the cross section roughly increasing with mass. The net result is a slight shift in the A$\\sim$195 abundance peak. Despite the ability to predict a more reasonable abundance of the A$\\sim$190 nuclei, there is still a discrepancy between the predicted abundance distribution in the rare earth region and that of the solar system. The abundances of the rare earth elements (A$\\sim$165) are predicted to be lower in abundance by about an order of magnitude than the A$\\sim$130 peak, in fair agreement with that of the solar system. This corresponds to the argument of the authors of reference \\cite{surman97}, who state that the rare-earth region is a robust feature of any dynamical calculation including post-production of the r-process progenitors. However, nuclei with 130$<$A$<$160 are overproduced slightly, removing the effect of the rare earth region being manifest as a peak, hence the appearance of the abundance distributions in the figure. It is obvious that no shell quenching has been employed in this preliminary model, as can be seen by the dip in the A$\\sim$180 abundance. While the abundance of the A$\\sim$180 nuclei relative to the A$\\sim$130 peak is similar to that of the solar system, the width of this dip is greater than that of the solar system abundance distribution. It has been mentioned \\cite{kratz93} that the underproduction of nuclei in this region might vanish if the quenching of shell closures for the very neutron-rich nuclei in this mass region is properly included. Because of the large number of nuclei involved, the improved semi-gross theory has been used to globally calculate decay rates. It is understood that the initial model represents a first attempt to gauge the effects of nuclear $\\beta$-decays in the hot environment of the r-process, and further study is warranted. In particular, a more accurate global calculation of decay rates is desired. Currently, the measurements of $\\beta$-decay rates are limited to either ground-state nuclei or long-lived isomers. However, with the advent of large neutron flux devices\\cite{petrasso96}, the measurement of the GT strength functions of nuclei in excited states from may become feasible within the next decade, allowing for experimental confirmation of the transition strengths of excited-state nuclei. \\ack The authors wish to acknowledge the helpful insight provided by K. Takahashi and comments by B.A. Brown. This work was funded by NSF grants PHY 9901241 and PHY9905241 and by the WMU Faculty Research and Creative Activities Support Fund (FRACASF) grant \\#06-005." }, "0808/0808.3786_arXiv.txt": { "abstract": "In continuation of previous work, numerical results are presented, concerning relativistically counter-streaming plasmas. Here, the relativistic mixed mode instability evolves through, and beyond, the linear saturation -- well into the nonlinear regime. Besides confirming earlier findings, that wave power initially peaks on the mixed mode branch, it is observed that, during late time evolution wave power is transferred to other wave numbers. It is argued that the isotropization of power in wavenumber space may be a consequence of weak turbulence. Further, some modifications to the ideal weak turbulence limit is observed. Development of almost isotropic predominantly electrostatic -- partially electromagnetic -- turbulent spectra holds relevance when considering the spectral emission signatures of the plasma, namely bremsstrahlung, respectively magneto-bremsstrahlung (synchrotron radiation and jitter radiation) from relativistic shocks in astrophysical jets and shocks from gamma-ray bursts and active galactic nuclei. ", "introduction": " ", "conclusions": "" }, "0808/0808.3924_arXiv.txt": { "abstract": "In the study of stars, the high energy domain occupies a place of choice, since it is the only one able to directly probe the most violent phenomena: indeed, young pre-main sequence objects, hot massive stars, or X-ray binaries are best revealed in X-rays. However, previously available X-ray observatories often provided only crude information on individual objects in the Magellanic Clouds. The advent of the highly efficient X-ray facilities XMM-Newton and Chandra has now dramatically increased the sensitivity and the spatial resolution available to X-ray astronomers, thus enabling a fairly easy determination of the properties of individual sources in the LMC. ", "introduction": "In the range of astronomical tools, X-rays are of particular interest. Indeed, the high-energy domain unveils the most energetic phenomena taking place in our Universe. Such processes are usually difficult to perceive at other wavelengths, though they provide important constraints for astrophysics.\\\\ In this context, the MCs are targets of choice. Their advantages are multiple: the known and small distance, together with the small angular size, small inclination, different metallicities, and recent star formation episodes are here as crucial as at other wavelengths. In addition, the low obscuration towards the MCs renders soft X-ray observations much easier while the numerous available data taken at other wavelengths ensure a correct, global analysis of the LMC X-ray sources.\\\\ X-rays associated with the LMC were first detected 40 years ago by a rocket experiment (\\cite{mar69}): the source appeared extended and soft, with a total luminosity estimated to 4$\\times 10^{38}$\\,erg\\,s$^{-1}$. It did not take long to distinguish a few individual sources in this X-ray emission, nicknamed LMC\\,X-1 to 6, thanks to the joint effort of the satellites Uhuru, Copenicus, OSO-7, and Ariel V (\\cite{leo71}, \\cite{rap74}, \\cite{mar75}, \\cite{gri77}). In the following decade, the Einstein observatory increased the number of known X-ray sources in the direction of the LMC to about a hundred (\\cite{lon81}, \\cite{wan91}). Finally, a sensitive survey undertaken by ROSAT provided another ten-fold increase in the total source number (\\cite[Haberl \\& Pietsch 1999a, Sasaki \\etal\\ 2000]{hab99a,sas00}).\\\\ However, only half of the detected X-ray sources truly belonged to the LMC and an even smaller fraction appeared to be associated with LMC stars. For example, of the 758 ROSAT-PSPC sources, only 144 were identified at first (\\cite[Haberl \\& Pietsch 1999a]{hab99a}): 15 as background AGNs or galaxies behind the LMC, 57 as foreground Galactic stars, 46 as SNRs and SNRs candidates, 17 as X-ray binaries (XRBs) and candidates, 9 as Supersoft sources (SSSs) and candidates. Using the observed X-ray properties (especially the hardness ratios), \\cite{hab99a} further proposed additional identifications (3 AGNs, 27 foreground stars, 9 SNRs and 3 SSSs) which yields a fraction of 20\\% of X-ray sources associated with LMC stars. Similar results were obtained with the ROSAT-HRI (\\cite[Sasaki \\etal\\ 2000]{sas00}): 397 detections among which 138 in common with the PSPC and 115 identified (10 AGNs, 52 foreground stars, 33 SNRs, 12 XRBs, 5 SSSs, and 3 hard sources which could either be AGNs or XRBs). If one considers the variable X-ray sources, the contamination by non-LMC objects is smaller. For the PSPC survey, the proposed identification of the 27 variable X-ray sources is 12 XRBs, 5 SSSs, 9 foreground stars and 1 Seyfert galaxy (\\cite[Haberl \\& Pietsch 1999b]{hab99b}); for the HRI survey, 26 variable sources were detected among which 8 XRBs, 4 SSSs, 6 foreground stars, 2 AGNs and 1 nova (\\cite[Sasaki \\etal\\ 2000]{sas00}): the fraction of LMC stellar objects among variable X-ray sources is thus 60\\%.\\\\ The current facilities possess much higher sensitivities and spatial/spectral resolution but unfortunately they also have a smaller field-of-view. Its non-zero extension penalizes the LMC, especially in comparison with the SMC ($\\sim$59 sq. deg. vs. $\\sim$18). This explains why, up to now, no full survey of the LMC has been performed with \\xmm\\ or \\ch. Nevertheless, smaller fields have been observed and it should be underlined that, though its coverage is patchy, the 2XMM catalog currently lists 5421 entries in the area of the ROSAT surveys! In addition, \\cite{hab03} reported the analysis of one deep \\xmm\\ observation of a northern region of the LMC (on the rim of the supergiant shell LMC4). While ROSAT had detected 34 sources in this field, \\xmm\\ data reveal 150 objects (detection limit 6$\\times 10^{32}$\\,erg\\,s$^{-1}$). In a selection of 20 bright or peculiar sources, the majority (10) are AGNs, but there are also 3 foreground stars, 2 SNRs, 4 HMXBs, and 1 SSSs \\footnote{This source was not confirmed by \\cite{kah08}.}. In addition, \\cite{sht05} analyzed 23 \\xmm\\ archival observations covering 3.8 sq. degrees of the LMC. With a detection limit of 3$\\times 10^{33}$\\,erg\\,s$^{-1}$, they detected 460 sources in the 2--8\\,keV band, in vast majority AGNs, and focused on 9 good XRBs candidates and 19 possible XRBs (see below for more details). Finally, a sensitive survey of the LMC in the hard X-ray domain (15\\,keV--10\\,MeV) has been performed with {\\sc Integral} (\\cite{got06}). Only a few sources have been detected: the X-ray binaries LMC\\,X-1, LMC\\,X-4, and PSR B0540--69, as well as two hard sources which might correspond to LMC binaries. These encouraging first results in both the soft and hard X-ray domains enlight the detection potential of the sensitive observatories of the current generation. \\\\ ", "conclusions": "" }, "0808/0808.1921_arXiv.txt": { "abstract": "We present cosmological constraints arising from the first measurement of the radial (line-of-sight) baryon acoustic oscillations (BAO) scale in the large scale structure traced by the galaxy distribution. Here we use these radial BAO measurements at $z=0.24$ and $z=0.43$ to derive new constraints on dark energy and its equation of state for a flat universe, without any other assumptions on the cosmological model: $w = -1.14 \\pm 0.39$ (assumed constant), $\\Omega_m = 0.24^{+0.06}_{-0.05}$. If we drop the assumption of flatness and include previous cosmic microwave background and supernova data, we find $w = -0.974 \\pm 0.058$, $\\Omega_m = 0.271 \\pm 0.015$, and $\\Omega_k = -0.002 \\pm 0.006$, in good agreement with a flat cold dark matter cosmology with a cosmological consant. To our knowledge, these are the most stringent constraints on these parameters to date under our stated assumptions. ", "introduction": " ", "conclusions": "" }, "0808/0808.0320_arXiv.txt": { "abstract": "We analyse intermediate-resolution VLT FLAMES/Giraffe spectra of six ultra-compact dwarf (UCD) galaxies in the Fornax cluster. We obtained velocity dispersions and stellar population properties by full spectral fitting against {\\sc pegase.hr} models. Objects span a large range of metallicities (--0.95 to --0.23~dex), 4 of them are older than 8~Gyr. Comparison of the stellar and dynamical masses suggests that UCDs have little dark matter at best. For one object, UCD3, the Salpeter initial mass function (IMF) results in the stellar mass significantly exceeding the dynamical one, whereas for the Kroupa IMF the values coincide. Although, this object may have peculiar dynamics or/and stellar populations, the Kroupa IMF seems more realistic. We find that UCDs lie well above the metallicity--luminosity relation of early-type galaxies. The same behaviour is demonstrated by some of the massive Milky Way globular clusters, known to contain composite stellar populations. Our results support two following UCD formation scenarii: (1) tidal stripping of nucleated dwarf elliptical galaxies; (2) formation of tidal superclusters in galaxy mergers. We also discuss some of the alternative channels of the UCD formation binding them to globular clusters. ", "introduction": "A new class of compact dwarf galaxies, called UCDs, was discovered a decade ago in the Fornax cluster \\citep{Hilker+99,Drinkwater+00,PDGJ01}. These objects, being brighter and much larger than globular clusters (GCs), but still far below luminosities and sizes of both dwarf elliptical (dE) and compact elliptical (cE) galaxies, fill an empty region on the Fundamental Plane \\citep{DD87}. Their origin still remains a matter of debate; several alternatives are considered: (1) UCDs are the result of the evolution of primordial density fluctuations \\citep{PDGJ01}; (2) they have been formed through mergers of GCs or simply represent the extreme high-luminosity end of the GC luminosity function \\citep{MHI02}; (3) UCDs are nuclei of tidally stripped (``threshed'') nucleated dE (dE,N) galaxies \\citep{BCDS03} or dE,Ns with very low surface brightness; (4) UCDs are created as tidal superclusters during major mergers of galaxies \\citep{FK05,KJB06}. Mass-to-light ratios of UCDs vary quite significantly \\citep{Drinkwater+03,Hasegan+05,Hilker+07}, suggesting the presence of dark matter in some of them. \\cite{Hasegan+05} propose to use $M/L$ ratio (i.e. presence of dark matter) as a criterion to distinguish between ``UCD galaxies'' and massive GCs. Developing this idea, we conclude that the presence of dark matter in a compact stellar system rejects two formation scenarii -- in this case UCDs cannot be GCs neither they can be created as tidal superclusters during galaxy mergers. On the other hand, if the stellar population is not old and metal-poor, the primordial density fluctuation scenario will become implausible, leaving the only channel of UCD formation to be the tidal stripping of dE,Ns. Stellar population analysis may also help to choose the formation scenario. Presently published data \\citep{MHIJ06,EGDH07} based on the analysis of absorption line strengths (Lick indices, \\citealp{WFGB94}) suggest that UCDs are old and rather metal-poor. In this paper, we present stellar population parameters for 6 Fornax cluster UCDs, compare them with dE,N nuclei, and derive stellar masses to check for the presence of dark matter assuming different stellar IMFs. ", "conclusions": "\\subsection{Comparison of Stellar and Dynamical Masses} Given the stellar population parameters and luminosities, we derive stellar masses of UCDs in our sample. The stellar mass estimates, computed from the mass-to-light ratios provided by {\\sc pegase.2} \\citep{FR97} for Salpeter and Kroupa et al. IMFs are given in Table~\\ref{tabmlcomp}. In the fourth column we provide the corrected dynamical masses, derived by re-normalising the values of \\cite{Hilker+07} by our velocity dispersion measurements ($M_{\\rm{d, corr}} = M_{\\rm{d}} (\\sigma / \\sigma_{\\rm{lit}})^2$. The fifth and sixth column contain the dark matter fractions estimated from Salpeter and Kroupa SSPs and corrected dynamical masses. For UCD1, 2, 4, and 5 the Salpeter SSP stellar masses are consistent with the dynamical ones within uncertainties, although in average the stellar masses tend to be lower. The Kroupa et al. IMF decreases them more resulting in a 40--50 per cent (65 for UCD5) upper limits of the dark matter content. However, for UCD3 the stellar mass derived using the Salpeter IMF becomes significantly larger than the dynamical estimate (note negative dark matter fraction), whereas Kroupa IMF provides almost a perfect match between the two, suggesting zero dark matter content. Therefore, if we assume no object-to-object IMF variation, the observations are more in favour of the Kroupa et al. IMF. There is a possibility that the dynamical model of UCD3 used by \\cite{Hilker+07} was not correct (for example, (1) the outer component of the UCD3 is not spherically symmetric or (2) there is significant rotation not taken into account, or (3) velocity dispersions are anisotropic), which may lead to an underestimated dynamical mass. At the same time, we cannot exclude that SSP models do not represent well the spectrum (e.g. the object contains a metal-rich sub-population, which is not properly modeled). In this case the stellar mass may be overestimated. Given the large uncertainties of stellar population parameters and, consequently, stellar mass-to-light ratios, we cannot give a decisive answer on a question \\emph{``Is there dark matter in UCDs?''} However, the main conclusion we draw is that \\emph{UCDs are not dark matter dominated objects} and at present level of detection, the dark matter content can be explained by uncertainties of the measurements and looseness of the models used to derive dynamical and stellar masses. \\begin{table} \\caption{Comparison of stellar masses of 6 UCDs for Salpeter (2) and Kroupa (3) IMFs; and the dynamical masses for 5 objects (5) from Hilker et al. (2007) corrected using our velocity dispersion estimations, (6)--(7) the dark matter content (per cent) for the Salpeter and Kroupa et al. IMF.\\label{tabmlcomp}} \\begin{tabular}{cccccc} \\hline $n$ & $M_{*{\\rm{Salp.}}}$ & $M_{*{\\rm{Kroupa}}}$ & $M_{\\rm{d, corr}}$ & DM$_{\\rm{S.}}$ & DM$_{\\rm{K.}}$ \\\\ & 10$^7 M_{\\odot}$ & 10$^7 M_{\\odot}$ & 10$^7 M_{\\odot}$ & \\% & \\% \\\\ \\hline 1 & 2.9$\\pm$0.7 & 2.1$\\pm$0.5 & 3.7$\\pm$0.5 & 20 & 45 \\\\ 2 & 2.3$\\pm$0.8 & 1.4$\\pm$0.5 & 2.4$\\pm$0.3 & 0 & 40 \\\\ 3 & 17.5$\\pm$3.0 & 12.8$\\pm$2.2 & 12.0$\\pm$2.4 &--45 & 0 \\\\ 4 & 2.9$\\pm$0.8 & 1.9$\\pm$0.6 & 4.0$\\pm$1.0 & 30 & 50 \\\\ 5 & 0.9$\\pm$0.2 & 0.5$\\pm$0.1 & 1.4$\\pm$0.5 & 35 & 65 \\\\ 6 & 2.3$\\pm$0.4 & 1.6$\\pm$0.4 & & & \\\\ \\hline \\end{tabular} \\end{table} \\subsection{Metallicity-$M_B$ and metallicity-$\\sigma$ relations} \\begin{figure} \\includegraphics[width=0.95\\hsize]{UCD_A496_sigma_met.ps}\\\\ \\includegraphics[width=0.95\\hsize]{UCD_A496_MB_met.ps} \\caption{Metallicity-velocity dispersion (top) and metallicity-luminosity (bottom) relations for early-type galaxies, UCDs, and GCs. The data sources are described in the text. Outlined crosses represent Galactic GCs with multi-component stellar populations (see Section 4.2 for details). \\label{figSLZ}} \\end{figure} In Fig.~\\ref{figSLZ} we plot metallicities versus velocity dispersions and luminosities of the 6 Fornax cluster UCDs and 2 dE,N nuclei from our sample. They are compared to: (a) Milky Way GCs from \\cite{MvdM05}, outlined crosses show GCs with direct evidences of multiple stellar populations revealed by the analysis of colour-magnitude diagrams \\citep{Piotto08}; (b) Local Group dEs and dwarf spheroidals (dSph) from \\cite{Mateo98}; (c) Abell~496 low-luminosity early-type galaxies from \\cite{Chil08A496}; (d) a sample of intermediate-luminosity and bright early-type galaxies from \\cite{SGCG06a,SGCG06c}; two compact stellar systems in the Virgo cluster with the spectra available in SDSS DR6 \\citep{SDSS_DR6}: transitional UCD/cE ``M59cO'' \\citep{CM08} and VUCD~7, where the data have been processed exactly in the same way as for M59cO. The aperture size for the Abell~496 dE/dS0 is around 0.8~kpc, so the dE nuclei do not dominate the light, therefore metallicities and velocity dispersions should be closer to the global than to the central values, given flat velocity dispersion profiles usually observed in dEs \\citep{SP02, GGvdM03, vZSH04}. In the metallicity-luminosity relation (Fig.~\\ref{figSLZ}, bottom panel) there is a continuous sequence $Z \\varpropto L_B^{0.45}$, spanning over 6 orders of magnitude in luminosity, formed by early-type galaxies: from the faintest dSph on the left to the brightest cluster ellipticals on the right. UCD galaxies lie significantly above (0.7--1.0~dex) this sequence, compared to the brightest Local Volume dSph's, having similar luminosities (see \\citealp{Mateo98} and references therein). In this sense, UCDs are similar to cE galaxies (e.g. \\citealp{Chilingarian+07}) having high metallicities for their luminosities, which probably represent end-products of the galaxy tidal threshing \\citep{BCD01}. At the same time, on the metallicity-velocity dispersion plot (Fig.~\\ref{figSLZ}, top panel) the loci of UCDs practically coincide with those of dEs. We consider this as an argument for the scenario of tidal threshing of dE,Ns as a way to create UCDs. In this case, a velocity dispersion of a compact nucleus will not change very strongly, while the total luminosity of a progenitor will drop by several magnitudes. For comparison with massive GCs \\citep{MvdM05} we chose a subsample of Galactic GCs with available measurements of velocity dispersions. Many GCs follow the behaviour of early-type galaxies. The three strongest outliers, namely NGC~104, NGC~6388, and NGC~6441, similarly to UCDs, reside significantly above the sequence of early-type galaxies on the metallicity -- luminosity plot (Fig~\\ref{figSLZ}). It is remarkable that the latter two exhibit direct evidences of multiple stellar populations \\citep{Piotto08}. Among the three other GCs (NGC~1851, NGC~2808, $\\omega$~Cen) demonstrating composite stellar populations only $\\omega$~Cen follows exactly the trend defined by early-type galaxies. We notice, that the third strongest outlier, 47~Tuc (NGC~104), being at least as massive as NGC~6388 does not have an evident double main sequence \\citep{Piotto08}. In the frame of the tidal stripping scenario, we can also propose an explanation for the large spread of UCD metallicities ($-$0.25 to $-$0.93) on the [Fe/H]~vs.~$\\sigma$ plot. It may be a superposition of the two factors: (1) the relatively high spread of metallicities of dE progenitors of UCDs due to their own environmentally-driven evolution (see discussion in \\citealp{Chil08A496}); (2) different conditions during the tidal stripping, which may lead to some changes of the velocity dispersion values compared to the progenitors. \\subsection{Comparison of dE,N nuclei and UCDs} In Fig.~\\ref{figtZ} we compare ages and metallicities of the 6 Fornax cluster UCDs with nuclei of dE,Ns in the Fornax (2 objects, this study) and Virgo (26 objects) clusters, transitional cE/UCD object M59cO \\citep{CM08} and VUCD7, another UCD in the Virgo cluster. Stellar population parameters for 22 Virgo cluster galaxies (shown in red), as well as for VUCD7 and M59cO are obtained by analysing SDSS DR6 spectra. For the four remaining Virgo dE,Ns shown in light blue we used the results based on the 3D spectroscopic observations presented in \\cite{CPSA07,CSAP07}: diamonds with error-bars correspond to the ages and metallicities of the nuclei and the blue vectors point to the parameters of the ``main bodies'' of the galaxies. Apart from the 2 intermediate age objects (UCD4 and UCD5), all UCDs are old, at the same time spanning a large range of metallicities. Most of the dE,Ns exhibit considerably younger stellar populations than UCDs. However, there is a number of old dE,N nuclei (including FCC~182) with ages comparable to those of UCDs. A scenario, assuming that dE,N nuclei are results of repeated or extended star formation episodes in the dE centres, leading to metal enrichment, can explain the observed quite high metallicities of dE,N nuclei. \\begin{figure} \\includegraphics[width=0.95\\hsize]{agemet_ucd_new.ps} \\caption{Comparison of ages and metallicities of UCDs with a sample of Virgo cluster dE,N galaxies from SDSS.\\label{figtZ}} \\end{figure} \\subsection{Origin of UCD galaxies} The stellar population parameters obtained by our SSP fitting, namely metallicities higher than $-$1.0 dex, allow us to exclude the scenario of evolving primordial density fluctuations \\citep{PDGJ01}, because one would expect much lower metallicities for objects at this mass range. Low dark matter content leaves space for all remaining channels of the UCD formation: \\cite{BCDS03} showed that in case of dE stripping (``threshing'') the progenitor's nucleus must not be dark matter dominated; the two other alternatives, GC merging and formation of UCDs as tidal superclusters, assume zero dark matter content. However, the scenario of merging GCs \\citep{MHI02} fails to explain why we do not observe metal-poor UCDs. It is known that GCs exhibit a dichotomy in the metallicity distribution (e.g. review by \\citealp{BS06}), but observed UCDs correspond only to metal-rich GCs. Why don't metal-poor GCs merge? Moreover, in case of a merger of metal-poor and metal-rich GCs of the same mass, the resulting luminosity-weighted metallicity in our wavelength range will be lower than the mean value, because metal-poor stellar populations have lower $M/L$ ratios than metal-rich ones. Composite stellar populations observed in massive Galactic GCs \\citep{Piotto08} comprise two to four SSPs, sometimes significantly different in metallicities, which is evident from deep colour-magnitude diagrams. These objects look good candidates for the GC merger scenario. In addition, we do see metal-poor ``composite'' GCs (NGC~1851, NGC~2808, $\\omega$~Cen). Indeed, they are an order of magnitude (except $\\omega$~Cen, where at least four SSPs are evident) fainter than the UCDs we are discussing here. On the statistical basis, UCDs are too frequent to be representatives of the high-luminosity end of the GC luminosity function (GCLF). After the initial discussion in \\cite{MHI02} a significant number of UCDs was discovered. The extrapolation of the GCLF (see e.g. \\citealp{ML07}) towards bright objects results in the statistical over-population of $M_V < -11$ objects. Although, the exact value depends on the adopted GCLF parameters and representation (i.e. Gaussian or $t_5$), we consider this fact important, making this channel of UCD formation scarcely probable. The old ages of most UCDs, compared to dE,N nuclei, suggest that if we consider the dE,N tidal stripping scenario, it must have happened a long time ago. However, there is a difficulty in explaining the formation of the most metal-rich UCDs, because 8--10~Gyr ago dE,N nuclei must have been less metal rich than presently observed. A possible explanation is a tidal stripping of the most massive dE representatives (dE/E transitional objects) such as IC~3653 or FCC~182. Another possibility is to create them as stellar superclusters (Fellhauer \\& Kroupa 2005) during interactions of massive galaxies, in this case metal-rich population formed from the metal-rich gas of the progenitors will be observed in the UCDs. Both scenarii are compatible with low dark matter content. A possible diagnosis is to measure $\\alpha$/Fe abundance ratios: populations formed in a short and intense star formation episode will be $\\alpha$-overabundant (e.g. \\citealp{Matteucci94}). With the present low S/N UCD data we are not able to carry out this test. Finally, we are left with the two alternatives of UCD formation: UCDs with low metallicities ($[\\mbox{Fe/H}] < -0.5$~dex) are in favour of dE,N tidal stripping, while tidally created superclusters better explain metal-rich UCDs. At present we cannot exclude the diversity of the UCD origin suggested by \\cite{MHIJ06}." }, "0808/0808.0044_arXiv.txt": { "abstract": "We consider the problem of optimal weighting of tracers of structure for the purpose of constraining the non-Gaussianity parameter $\\fnl$. We work within the Fisher matrix formalism expanded around fiducial model with $\\fnl=0$ and make several simplifying assumptions. By slicing a general sample into infinitely many samples with different biases, we derive the analytic expression for the relevant Fisher matrix element. We next consider weighting schemes that construct two effective samples from a single sample of tracers with a continuously varying bias. We show that a particularly simple ansatz for weighting functions can recover all information about $\\fnl$ in the initial sample that is recoverable using a given bias observable and that simple division into two equal samples is considerably suboptimal when sampling of modes is good, but only marginally suboptimal in the limit where Poisson errors dominate. ", "introduction": "\\label{sec:introduction} The currently most attractive theory for the emergence of structure in the Universe is inflation \\cite{Starobinsky:1979ty,1981PhRvD..23..347G,1982PhLB..108..389L,1982PhRvL..48.1220A}. It is generically successful at diluting the primordial defects to undetectable densities and predicts a nearly-flat universe with nearly scale invariant spectrum of primordial fluctuations that are normally distributed and extend to scales larger than horizon \\cite{1981JETPL..33..532M,1982PhLB..115..295H,1982PhRvL..49.1110G,1982PhLB..117..175S,1983PhRvD..28..679B}. To understand details of the inflation, on must look at detailed predictions of different models. Non-Gaussianity of the primordial curvature perturbations, i.e small departures from the normal distribution of fluctuations is one aspect in which models of inflation differ. Recently, non-Gaussianity of the local $\\fnl$ type has received a renewed attention. This type of non-Gaussianity is characterised by a quadratic correction to the potential \\citep{1990PhRvD..42.3936S,1994ApJ...430..447G,2000MNRAS.313..141V,2001PhRvD..63f3002K}: \\begin{equation} \\Phi = \\phi + \\fnl \\left( \\phi^2 - \\left<\\phi^2 \\right> \\right), \\label{fnl} \\end{equation} where $\\phi$ is the primordial potential assumed to be a Gaussian random field and $\\fnl$ describes the amplitude of the correction during the matter domination era. There are two main reasons for this renewed interest. First, there is a hint of a detection in the cosmic microwave data \\cite{2008PhRvL.100r1301Y} and several non-detections \\cite{2007JCAP...03..005C,2008arXiv0803.0547K,2008arXiv0802.3677H}. Second, a new method for its detection has been recently proposed in \\cite{Dalal:2007cu}. This method uses biased tracers of structure for which it can be shown that local-type of non-Gaussianity leads to a very particular scale-dependence of the bias \\begin{equation} \\Delta b = \\fnl (b-1) u (k), \\label{eq:4} \\end{equation} where $\\Delta b$ is the bias induced by non-Gaussianity, $b$ is the tracer's intrinsic bias and $u$ is given by \\begin{equation} u(k) = \\frac{3 \\delta_c \\Omega_m H_0^2}{c^2 k^2 T(k) D(z)}, \\label{eq:3} \\end{equation} where $T(k)$ is the matter transfer function normalised to unity at $k=0$, $D(z)$ is the growth function normalised to $(1+z)^{-1}$ in the matter era, $\\delta_c=1.68$ is the linear over-density at collapse for the spherical collapse model and other symbols have their usual meaning. Note that $\\Delta b$ becomes significant only at large scales, where non-linearities and scale-dependent bias are expected to be small and therefore offers a surprisingly clean probe of non-Gaussianity. This equation has been re-derived, scrutinised and better understood in the subsequent work \\cite{2008ApJ...677L..77M,2008arXiv0805.3580S,2008arXiv0806.1046A,2008arXiv0806.1061M}. A first application of this method to the real data using a wide variety of tracers of large scale has recently shown the promise of this method \\cite{2008arXiv0805.3580S,2008arXiv0806.1046A}. The derived constraints are already competitive with those coming from the cosmic microwave background. In that work, the constraints were derived by comparing the power spectrum of the distribution of tracers with those predicted by the theory. At largest scales, where the effect coming from the non-Gaussianity is the largest, the method suffers from the sample variance. In other words, the finite number of large-scale modes in any survey severely limits our ability to measure the power spectrum. Recently, Seljak has suggested a method of circumventing this limitation \\citep{2008arXiv0807.1770S}. This method essentially considers two differently biased tracers that sample the same volume. The ratio of amplitudes of a single mode for the two tracers will give the ratio of the two biases $(b_1+\\Delta b_1(\\fnl))/(b_2+\\Delta b_2(\\fnl))$, but the amplitude of the primordial mode cancels out. One thus measures the auto correlation power spectra of the two tracers in the same volume. By taking the ratio of these two spectra, one can put a constraint on the value of $\\fnl$, which is independent on the primordial field and thus unaffected by the sample variance. The biases $b_1$ and $b_2$ can be derived from the amplitude of small scale fluctuations, where sampling variance is not a problem and hence, one extremely well measured large-scale mode is in principle enough to constrain $\\fnl$. A more robust technique would be to assume nothing about the matter power spectrum and derive limits on $\\fnl$ from limits on the scale dependence of ratio of $b+\\Delta b$. This would protect measurements of $\\fnl$ from systematics arising from, for example, massive neutrinos. In practice, one rarely has two distinct samples with a well-defined bias. In this work, we extended the analysis by considering a single tracer of the underlying field that spans a range of biases and attempt to answer the question of how to optimally analyse such tracer. The approach we take is to create two effective samples and to optimally weight the tracer's constituents. ", "conclusions": "\\label{sec:conclusion} In this paper we have analysed the problem of optimal weighting of biased tracers of structure with the goal of extracting maximum information about the non-Gaussianity parameter $\\fnl$. We have derived the minimum error on $\\fnl$ by considering slices that are infinitely thin in bias. We have shown that a simple weighting scheme of Equations \\eqref{eq:1} and \\eqref{eq:2} obtains the same constraining power. General division of the full sample into two subsamples can be considerably sub-optimal even when mass at which the samples are divided is carefully chosen. The optimal weighting scheme of Equations \\eqref{eq:1} and \\eqref{eq:2} is surprisingly simple. In fact, the product $P\\bar{n}$ does not come into weighting at all - this is a lucky coincidence, which allows us to use the same optimal weighting for every mode, rather to optimize weighting around some fiducial wave-vector. The result in this paper is subject to the assumptions outlined in the Section \\ref{sec:appr-limit-this} of this paper. If these assumptions are violated, the weighing is sub-optimal, but probably nevertheless beneficiary. Since any division into two samples by e.g. an absolute magnitude cut requires some knowledge of bias, the implementation of the scheme proposed in this paper is likely to be very simple. How can this be put in practice? In this work we have used halo mass $M$ as a proxy for the bias. However, our analysis is completely general and one can replace the host halo mass with any variable that is monotonically linked to the bias. For example, one could take luminous red galaxies (LRGs) and determine their bias by splitting the entire sample into several subsamples in different luminosity bins and the constrain a smooth function $b(L)$, which describes the variation of galaxy bias with its luminosity, using modes which are not affected by the $\\fnl$. One would next construct two effective samples by optimally weighting the original sample using Equations \\eqref{eq:1} and \\eqref{eq:2} and replacing $b(M)$ with $b(L)$. In the next step, auto and cross-correlation power spectra of these two samples should be calculated, taking into account the Poisson error correlation between the two. At this step, one can use the cross-correlation spectra to check for the amount of stochasticity, which has been assumed to be negligible in this work. Finally, $\\fnl$ should be constrained using these power spectra as input." }, "0808/0808.2870_arXiv.txt": { "abstract": "We report new precision measurements of the properties of our Galaxy's supermassive black hole. Based on astrometric (1995-2007) and radial velocity (2000-2007) measurements from the W. M. Keck 10-meter telescopes, a fully unconstrained Keplerian orbit for the short period star S0-2 provides values for the distance (R$_0$) of 8.0 $\\pm$ 0.6 kpc, the enclosed mass (M$_{bh}$) of 4.1 $\\pm$ 0.6 $\\times$ $10^6 M_{\\odot}$, and the black hole's radial velocity, which is consistent with zero with 30 km/s uncertainty. If the black hole is assumed to be at rest with respect to the Galaxy (e.g., has no massive companion to induce motion), we can further constrain the fit and obtain R$_0$ = 8.4 $\\pm$ 0.4 kpc and M$_{bh}$ = 4.5 $\\pm$ 0.4 $\\times$ $10^6 M_{\\odot}$. More complex models constrain the extended dark mass distribution to be less than 3-4 $\\times$ $10^5 M_{\\odot}$ within 0.01 pc, $\\sim$100x higher than predictions from stellar and stellar remnant models. For all models, we identify transient astrometric shifts from source confusion (up to 5x the astrometric error) and the assumptions regarding the black hole's radial motion as previously unrecognized limitations on orbital accuracy and the usefulness of fainter stars. Future astrometric and RV observations will remedy these effects. Our estimates of R$_0$ and the Galaxy's local rotation speed, which it is derived from combining R$_0$ with the apparent proper motion of Sgr A*, ($\\theta_0$ = 229 $\\pm$ 18 km s$^{-1}$), are compatible with measurements made using other methods. The increased black hole mass found in this study, compared to that determined using projected mass estimators, implies a longer period for the innermost stable orbit, longer resonant relaxation timescales for stars in the vicinity of the black hole and a better agreement with the M$_{bh}$-$\\sigma$ relation. ", "introduction": "\\label{sec:intro} Ever since the discovery of fast moving (v $>$ 1000 km s$^{-1}$) stars within 0.$\\tt''$3 (0.01 pc) of our Galaxy's central supermassive black hole (Eckart \\& Genzel 1997; Ghez et al. 1998), the prospect of using stellar orbits to make precision measurements of the black hole's mass (M$_{bh}$) and kinematics, the distance to the Galactic center (R$_0$) and, more ambitiously, to measure post-Newtonian effects has been anticipated (Jaroszynski 1998, 1999; Salim \\& Gould 1999; Fragile \\& Mathews 2000; Rubilar \\& Eckart 2001; Weinberg, Milosavlejic \\& Ghez 2005; Zucker \\& Alexander 2007; Kraniotis 2007; Will 2008). An accurate measurement of the Galaxy's central black hole mass is useful for putting the Milky Way in context with other galaxies through the apparent relationship between the mass of the central black hole and the velocity dispersion, $\\sigma$, of the host galaxy (e.g., Ferrarese \\& Merrit 2000; Gebhardt et al. 2000; Tremaine et al. 2002). It can also be used as a test of this scaling, as the Milky Way has the most convincing case for a supermassive black hole of any galaxy used to define this relationship. Accurate estimates of R$_0$ impact a wide range of issues associated with the mass and structure of the Milky Way, including possible constraints on the shape of the dark matter halo and the possibility that the Milky Way is a lopsided spiral (e.g., Reid 1993; Olling \\& Merrifield 2000; Majewski et al. 2006). Furthermore, if measured with sufficient accuracy ($\\sim$1\\%), the distance to the Galactic center could influence the calibration of standard candles, such as RR Lyrae stars, Cepheid variables and giants, used in establishing the extragalactic distance scale. In addition to estimates of $M_{bh}$ and R$_0$, precision measurements of stellar kinematics offer the exciting possibility of detecting deviations from a Keplerian orbit. This would allow an exploration of a possible cluster of stellar remnants surrounding the central black hole, suggested by Morris (1993), Miralda-Escud{\\'e} \\& Gould(2000), and Freitag et al. (2006). Estimates for the mass of the remnant cluster range from $10^4 - 10^5 M_{\\odot}$ within a few tenths of a parsec of the central black hole. Absence of such a remnant cluster would be interesting in view of the hypothesis that the inspiral of intermediate-mass black holes by dynamical friction could deplete any centrally concentrated cluster of remnants. Likewise, measurements of post-newtonian effects would provide a test of general relativity, and, ultimately, could probe the spin of the central black hole. Tremendous observational progress has been made over the last decade towards obtaining accurate estimates of the orbital parameters of the fast moving stars at the Galactic center. Patience alone permitted new astrometric measurements that yielded the first accelerations (Ghez et al. 2000; Eckart et al. 2002), which suggested that the orbital period of the best characterized star, S0-2, could be as short as 15 years. The passage of more time then led to full astrometric orbital solutions (Sch\\\"odel et al. 2002, 2003; Ghez et al. 2003, 2005a), which increased the implied dark mass densities by a factor of $10^4$ compared to earlier velocity dispersion work and thereby solidified the case for a supermassive black hole. The advent of adaptive optics enabled radial velocity measurements of these stars (Ghez et al. 2003), which permitted the first estimates of the distance to the Galactic center from stellar orbits (Eisenhauer et al. 2003, 2005). In this paper, we present new orbital models for S0-2. These provide the first estimates of the distance to the Galactic center and limits on the extended mass distribution based on data collected with the W. M. Keck telescopes. The ability to probe the properties of the Galaxy's central supermassive black hole has benefitted from several advancments since our previous report (Ghez et al. 2005). First, new astrometric and radial velocity measurements have been collected between 2004 and 2007, increasing the quantity of kinematic data available. Second, the majority of the new data was obtained with the laser guide star adaptive optics system at Keck, improving the quality of the measurements (Ghez et al. 2005b; Hornstein et al. 2007). These new data sets are presented in \\S\\ref{sec:obs}. Lastly, new data analysis has improved our ability to extract radial velocity estimates from past spectroscopic measurements, allowing us to extend the radial velocity curve back in time by two years, as described in \\S\\ref{sec:data_analysis}. The orbital analysis, described in \\S\\ref{sec:orbit}, identifies several sources of previously unrecognized biases and the implications of our results are discussed in \\S\\ref{sec:disc}. ", "conclusions": "" }, "0808/0808.2277_arXiv.txt": { "abstract": "{The quasi-Hilda comets (QHCs), being in unstable 3:2 Jovian mean motion resonance, are considered a major cause of temporary satellite capture (TSC) by Jupiter. Though the QHCs may be escaped Hilda asteroids, their origin and nature have not yet been studied in sufficient detail. Of particular interest are long TSCs/orbiters. Orbiters -- in which at least one full revolution about the planet is completed -- are rare astronomical events; only four have been known to occur in the last several decades. Every case has been associated with a QHC: 82P/Gehrels 3; 111P/Helin-Roman-Crockett; P/1996 R2 (Lagerkvist); and the possibly QHC-derived D/1993 F2 (Shoemaker-Levy 9, SL9). } {We focus on long TSC/orbiter events involving QHCs and Jupiter. Thus we survey the known QHCs, searching for further long TSCs/orbiters over the past century. } {First, we confirmed the long TSC/orbiter events of 82P, 111P, and 1996 R2 in order to test our method against previous work, applying a general $N$-body Newtonian code. We then used the same procedure to survey the remaining known QHCs and search for long TSC/orbiter events. } {We newly identified another long TSC/orbiter: 147P/Kushida-Muramatsu from 1949 May 14$^{+97 {\\rm days}}_{-106 {\\rm days}}$--1961 July 15. Our result is verified by integrations of 243 cloned orbits which take account of the present orbital uncertainty of this comet. The event involves an $L_{\\rm 2} \\to L_{\\rm 1}$ transition as with 82P and 1996 R2; this may represent a distinct subtype of TSCs from QHC derived ($L_{\\rm 1} \\to$) longer captures exemplified by 111P and (probably) SL9, though this classification is still only based on a small database of TSCs. } {This is the third long TSC and the fifth orbiter to be found, thus long TSC/orbiter events involving Jupiter have occurred once per decade. Two full revolutions about Jupiter were completed and the capture duration was $12.17^{+0.29}_{-0.27}$ years; both these numbers rank 147P as third among long TSC/orbiter events, behind SL9 and 111P. This study also confirms the importance of the QHC region as a dynamical route into and out of Jovian TSC, via the Hill's sphere. } ", "introduction": "Among all the asteroids that have been recorded up to the present in the asteroid database, e.g., the ``JPL Small-Body Database\" ({\\tt http://ssd.jpl.nasa.gov/}), a large number (more than 1000 including unnumbered objects) are known to populate the region of the 3:2 mean motion resonance (MMR) with Jupiter, in the outer main belt. These are the ``Hilda asteroids\" (Schubart \\cite{schubart68}, \\cite{schubart82}, \\cite{schubart}; Ip \\cite{ip}; Yoshikawa \\cite{yoshikawa}; Franklin et al. \\cite{franklin}; Nesvorn\\'y \\& Ferraz-Mello \\cite{nesvorny}). Their semimajor axes, $a$, concentrate in the range 3.7 AU $\\le a \\le$ 4.2 AU, at eccentricities $e \\le 0.3$, and inclinations $i \\le 20^\\circ$ (Zellner et al. \\cite{zellner}). This results in a range of the Tisserand parameter with respect to Jupiter, $T_{\\rm J}$, of $\\sim 2.90$--3.05, where $T_{\\rm J} = a_{\\rm J}/a + 2 \\sqrt {a/a_{\\rm J}(1 - e^2)} \\cos I$, with $a_{\\rm J}$ being the semimajor axis of Jupiter and $I$ the mutual inclination between the orbits. The critical arguments for the Hildas, $\\phi = 3\\lambda_{\\rm J}-2\\lambda-\\varpi$, librate about $0^\\circ$, being stable in the long-term ($\\lambda$ is mean longitude, $\\varpi$ is longitude of perihelion, and J indicates Jupiter). As regards physical properties, the low-albedo D- and P-types are more abundant in Hildas' surface colours than the small fraction of C-types (Dahlgren \\& Lagerkvist \\cite{dahlgren}; Dahlgren et al. \\cite{dahlgren97}; Gil-Hutton \\& Brunini \\cite{gil-hutton}; Licandro et al. \\cite{licandro}). The surface colour of D- and P-type asteroids, such as Hildas and Trojans in the outer main belt, corresponds well with that of cometary nuclei (Fitzsimmons et al. \\cite{fitzsimmons}; Jewitt \\cite{jewitt02}), which means that they are covered with a similar mineralogical surface, suggestive of a common origin. More than 50 Jovian irregular satellites are known at present (Jewitt \\& Haghighipour \\cite{jewitt}). \\'Cuk \\& Burns (\\cite{cuk}) pointed out that the progenitor of the main prograde cluster, the Himalia family, was plausibly derived from the Hildas long ago, if it was captured by a gas-drag assisted mechanism. Thus tracing the origin and nature of irregular satellites is very significant for studying the accretion processes in the early solar system. Satellite capture mechanics in the circular restricted three-body problem (CR3BP) or the $N$-body problem has often been investigated (e.g., H\\'enon \\cite{henon}; Huang \\& Innanen \\cite{huang}; Tanikawa \\cite{tanikawa}; Murison \\cite{murison}; Brunini et al. \\cite{brunini}; Nesvorn\\'y et al. \\cite{nesvorny03}, \\cite{nesvorny07}). Reflectance spectra of Jovian irregulars, being dominated by D- and C-types (Luu \\cite{luu}; Grav et al. \\cite{grav}), are comparable to those of Hildas. During the past half century, several Jupiter family comets (JFCs; cf.\\ Levison \\cite{levison}) have stayed in or near the Hilda zone, although being in unstable 3:2 MMR with Jupiter. Some of them have been transferred from outside to inside, \\textit{vice versa}, or from inside to inside of Jupiter's orbit by undergoing a temporary satellite capture (TSC) by Jupiter (e.g., Carusi \\& Valsecchi \\cite{carusi79}; Tancredi et al. \\cite{tancredi}). Such a JFC is called a ``quasi-Hilda comet\" (QHC) by Kres\\'ak (\\cite{kresak}), who identified three such objects: 39P/Oterma, 74P/Smirnova-Chernykh, and 82P/Gehrels 3. Di Sisto et al. (\\cite{di sisto}) integrated the motions of 500 fictitious Hildas for $\\sim 10^9$ years, and found that most of them escaped from the Hilda zone into the JFC population, i.e., left the 3:2 MMR and evolved quickly on to unstable orbits: such a chaotic diffusion from the Hilda zone has also been demonstrated by Nesvorn\\'y \\& Ferraz-Mello (\\cite{nesvorny}). In addition, large-scale collisional processes, such as the late heavy bombardment, might also release small bodies from the Hilda zone into JFCs, e.g., see Gil-Hutton \\& Brunini (\\cite{gil-hutton00}). Hence, some QHCs may indeed be such escaped Hildas themselves. Recently, Toth (\\cite{toth}) updated the QHC list, finding a total of 17 members (see Section 3). This includes bodies such as the renowned Comet D/1993 F2 (Shoemaker-Levy 9, SL9) that have undergone TSC by Jupiter and then disappeared after colliding with the planet. Surface spectroscopic (or colorimetric) measurements for QHCs have only been carried out for 82P (De Sanctis et al. \\cite{de sanctis}), the results of which also indicate a taxonomic D-type. Among the QHCs, 39P/Oterma was the first known to be temporarily captured by Jupiter, in 1936--1938 (Marsden \\cite{marsden}). However, this comet flew through the region near Jupiter over a rather short time, during which the comet did not complete a full revolution orbiting about the planet. In contrast, unlike 39P's ``fly-through\" capture, there is a different kind of TSC, in which at least one full revolution about the planet is completed; we deal with these in the present paper. Following Kary \\& Dones (\\cite{kary}) we call such objects ``orbiters\". These are often characterized by a long capture with very small perijove distance, usually lasting for $\\sim 10$ years or more. Not all orbiters become such long TSCs ($> 10$ yr), although of course they last longer than the fly-through type. SL9 is a representative case for both long TSCs and orbiters. This comet was pointed out to have possibly been QHC-derived before its tidal disruption on passing through perijove at less than 1.5 Jovian equatorial radii ($R_{\\rm J}$, where $R_{\\rm J}=71492.4$ km), i.e., within the Roche limit, in 1992 July and its subsequent collision with Jupiter in 1994 July (Nakano \\& Marsden \\cite{nakano93}; Sitarski \\cite{sitarski}; Benner \\& McKinnon \\cite{benner}). If it is QHC-derived, then SL9 is the only QHC so far that has been orbiting the planet as a TSC at the time of discovery (Benner \\& McKinnon \\cite{benner}). By numerically integrating SL9's pre-collision orbital motion, several studies showed that the TSC duration of SL9 lasted for 50 years or more, during which the comet completed more than 30 revolutions orbiting about Jupiter, making it nominally the longest known TSC (Carusi et al. \\cite{carusi}; Benner \\& McKinnon \\cite{benner}; Chodas \\& Yeomans \\cite{chodas}). However, according to Benner \\& McKinnon (\\cite{benner}), SL9 was the most chaotic known object in the solar system with an effective Lyapunov time of only $\\sim10$ years on its jovicentric orbit. Thus it is difficult to determine with certainty SL9's pre-capture orbit and its true TSC duration. Nevertheless, Benner \\& McKinnon's additional statistical analysis of distributions in $a$-$e$ space and $T_{\\rm J}$ values, based on back integrations of the various SL9 fragments, revealed a possible QHC origin of SL9. The work of Kary \\& Dones (\\cite{kary}) supports this possibility: they traced the motions of numerous fictitious JFCs for $\\sim 10^5$ years and found that half the SL9-like very long captures $>50$ years were due to QHCs. They estimated that the frequency of such a very long TSC (= ``long capture\" as designated by them) is extremely rare, only 0.02\\% of all the TSC events in their simulations. Interestingly, impacts on Jupiter are more frequent than the very long TSCs by a factor of 8--9. They also evaluated that long TSCs (= ``orbiters bound $> 10$ yr\" as designated by them) and orbiters are still rare events, at the respective levels of 0.8\\% ($\\supset$ very long TSCs) and 2\\% ($\\supset$ long TSCs) relative to all events, with about 98\\% being the short TSC type which contains the fly-through events. Carusi \\& Valsecchi (\\cite{carusi79}) had earlier confirmed the rarity of orbiter events, simulating motions of fictitious small bodies as well. In the jovicentric Keplerian system, a TSC (especially a long TSC/orbiter) occurs whenever a small body passes near one of the collinear libration points $L_{\\rm 1}$ or $L_{\\rm 2}$ in the CR3BP of the Sun-Jupiter-(third) body system with very low velocity, i.e., effectively becoming bound by Jupiter when it enters the Hill's region with near-zero velocity. After that, the bound small body revolves about Jupiter on an elliptical jovicentric orbit until it again passes through the region near either $L_{\\rm 1}$ or $L_{\\rm 2}$ and escapes from the Jovian system. Considering such a transition using newly developed dynamical systems techniques based on a Hamiltonian formulation in the CR3BP, Koon et al. (\\cite{koon}) and Howell et al. (\\cite{howell}) demonstrated that a TSC by Jupiter occurs when the small body passes through a region inside the invariant manifold structure related to periodic halo orbits around $L_{\\rm 1}$ or $L_{\\rm 2}$ in the Hill's region. The TSC (or its duration) is usually defined by the jovicentric Kepler energy, $E_{\\rm J}$, being negative, $E_{\\rm J} < 0$, with the additional condition that the bound small body must be within the jovicentric sphere of gravitational influence: Kary \\& Dones (\\cite{kary}) set its boundary at 3 Hill's sphere radii (=1.065 AU) of Jupiter. However, Howell et al. (\\cite{howell}) defined the TSC duration as the residence time in the Hill's region. The former generally lasts longer than the latter, and here we regard the former as the TSC duration. The dynamics involved in TSC is quite different from that of quasi-satellites in 1:1 libration with Jupiter far outside the Hill's region (Wiegert et al. \\cite{wiegert}; Kinoshita \\& Nakai \\cite{kinoshita}). Apart from SL9, only three orbiters have been known to occur. Every case has been associated with a QHC: 82P; 111P/Helin-Roman-Crockett; and, though it is not in Toth's (\\cite{toth}) QHC list, P/1996 R2 (Lagerkvist). These TSCs were found by Rickman (\\cite{rickman}), Tancredi et al. (\\cite{tancredi}), and Hahn \\& Lagerkvist (\\cite{hahn}), respectively, and are discussed further in Section 2. The occurrence of these few events during the last several decades is consistent with the rarity of long TSC/orbiter events suggested by Kary \\& Dones (\\cite{kary}). Although 74P encountered Jupiter at distances of 0.24 AU and 0.47 AU in 1955 October and 1963 September respectively, it was not bound to Jupiter (Rickman \\cite{rickman}; Carusi et al. \\cite{carusi85b}); this comet is, however, expected to experience a TSC by Jupiter in this century (Carusi et al. \\cite{carusi85b} and see also Section 4). There further exist some JFCs that encountered Jupiter more closely than several QHCs involved in TSC, e.g., 16P/Brooks 2; D/1770 L1 (Lexell); 81P/Wild 2; but they were not captured by the planet owing to their high-velocity encounters (Carusi et al. \\cite{carusi85a}; Emel'yanenko \\cite{emel'yanenko}). Therefore studying the origin and nature of QHCs is of great importance from various astronomical points of view mentioned above, especially as regards their origin and being possibly related to the Hildas and the Jovian irregular satellites. Such studies may provide unique knowledge and clues about formation processes in the early solar system. Here we focus on long TSC/orbiter events involving Jupiter and QHCs. First (Section 2), applying a general $N$-body Newtonian code, we reconfirmed the long TSC/orbiter events of 82P, 111P, and 1996 R2. Then (Section 3) we used the same procedure to search Toth's (\\cite{toth}) QHCs list for other objects that have become long TSCs/orbiters in the past century. Eventually, we successfully found another long TSC/orbiter, occurring in the mid-20th century, 147P/Kushida-Muramatsu. ", "conclusions": "On the basis of our investigations above, we have presented a newly identified TSC of a comet by Jupiter. This is in the rare, orbiter class of TSCs and involves 147P from 1949 May 14$^{+97 {\\rm days}}_{-106 {\\rm days}}$--1961 July 15. This is the third long TSC of $>10$ years and the fifth orbiter found, so that Jupiter's long TSCs/orbiters have occurred once per decade. The completion of two full revolutions about Jupiter and the capture duration of $12.17^{+0.29}_{-0.27}$ years rank 147P as third in both these numbers among known orbiters, behind SL9 and 111P. \\subsection{TSC Classification} Following Kary \\& Dones (\\cite{kary}), we classify the known TSCs as follows: 39P as fly-through, 82P and 1996 R2 as orbiters, 111P and 147P as long TSCs ($>10$ years), and SL9 as a very long TSC ($>50$ years). On the other hand, depending on the TSC characteristics, Howell et al. (\\cite{howell}) defined two TSC types: Type 1 as a 39P-like fly-through; Type 2 as a long lasting capture like 111P, where the comet experiences more than one close encounter with Jupiter while in the TSC region. Here we tentatively divide Type 2 into two subtypes: Type 2A as the $L_{\\rm 2} \\to L_{\\rm 1}$ transition such as 82P, 1996 R2 and 147P; Type 2B as the QHC-derived ($L_{\\rm 1} \\to$) longer capture than 2A, as with 111P and SL9. Further TSC classifications may be possible in the future, if different kinds of TSC are found. \\subsection{Future TSCs} We also surveyed all the known QHCs, integrating their orbital motions forward for 100 years to check for future long TSCs/orbiters. We found that 111P will undergo a long TSC/orbiter phase with 6 perijove passages, and minimum $E_{\\rm J} \\sim -3.19$, from 2068 April 20--2086 June 09 (duration $\\sim 18.14$ years). We identified a rather long capture of 82P with minimum $E_{\\rm J} \\sim -2.26$ though it is not an orbiter but instead follows a symmetric trajectory about the $x$-axis, from 2056 February 18--2064 July 26 (duration $\\sim 8.43$ years). There appear to be two TSCs for 74P if we take its initial parameters from the JPL Small-Body Database, from 2025 August 10--2031 May 29 (duration $\\sim 5.80$ years, minimum $E_{\\rm J} \\sim -0.74$) and 2081 January 29--2085 January 29 (duration $\\sim 4.00$ years, minimum $E_{\\rm J} \\sim -0.77$). Such future TSCs have already been predicted by Carusi et al. (\\cite{carusi85a}, \\cite{carusi85b}) for 74P and 82P and Tancredi et al. (\\cite{tancredi}) for 111P, and our simulations are almost identical in profile with those. Meanwhile, it is unlikely that 1996 R2 and 147P will encounter Jupiter as a long TSC/orbiter for the next 100 years from each initial epoch. No further long TSCs/orbiters were found among the remaining QHCs. \\subsection{Tidal Effects} Here we discuss the Jovian tidal force acting on 147P and other captured QHCs, around their perijove passages. Well-known cometary tidal splitting events have occurred twice: 16P in 1886 (Sekanina \\& Yeomans \\cite{sekanina}) and SL9, both passing perijove within the Roche limit for comets, $\\sim 2.7 R_{\\rm J}$. In contrast, the close encounter of 147P with Jupiter around 1952 August 26 (Table~\\ref{tbl:147PTSC}) was at a perijove distance of $14.61^{+2.94}_{-2.61} R_{\\rm J}$, comfortably outside the Roche limit. The radius, $R_{\\rm c}$, of 147P's nucleus was estimated at 0.21 km by Lamy et al. (\\cite{lamy}), the smallest among all their measured cometary nuclei. The Jovian tidal stress, $\\sigma_{\\rm T}$, acting on small bodies (QHCs here), is given by: \\begin{equation} \\sigma_{\\rm T} \\approx Gm_{\\rm J}\\rho R_{\\rm c}^2/r_{\\rm J}^3, \\end{equation} where $m_{\\rm J}$ is the mass of Jupiter. Compared to the $\\sigma_{\\rm T}$ acting on SL9 at perijove $< 1.5 R_{\\rm J}$ in 1992 (Scotti \\& Melosh \\cite{scotti}; Asphaug \\& Benz \\cite{asphaug94}, \\cite{asphaug}), the $\\sigma_{\\rm T}$ for 147P was extremely small, amounting to $< 0.005 \\%$ if we take the original, pre-encounter $R_{\\rm c}$ of SL9 as $\\sim 1$ km (Scotti \\& Melosh \\cite{scotti}; Asphaug \\& Benz \\cite{asphaug94}) and $\\rho$ of both the comets as being equal. Hence, although our treatment here is approximate, there seems no reason to believe that 147P was affected by Jovian tides. On the other hand, 82P, which passed its perijove at $3.01 R_{\\rm J}$ in 1970, just outside the cometary Roche limit, may have suffered some effect due to the Jovian tides, even though it was not broken up. Taking $R_{\\rm c}$ for 82P as 0.73 km (Lamy et al. \\cite{lamy}), and equal $\\rho$ as above, implies that the $\\sigma_{\\rm T}$ acting on 82P could be up to $7 \\%$ of that for SL9. If 82P is a friable and furthermore a loose rubble-pile object, the heating energy from continual Jovian tides might sufficiently affect the comet's nucleus structure so as to allow H$_2$O ice, if present, to sublimate even though 82P is beyond the usual heliocentric distance for H$_2$O sublimation. If our hypothetical scenario is true, such tidal heating effects could trigger cometary activity on 82P and produce an outburst. In this event, the coma size and consequent brightness would peak several days (or more) after the maximum tidal effects. Another intriguing point is whether or not other tidal effects, e.g., tidal distortion and tidal torques leading to rotation state changes (Scheeres et al.\\ \\cite{scheeres}), are detectable in the light-curve observations of the captured QHCs. In the physical database of cometary nuclei by Lamy et al. (\\cite{lamy}), we can notice a rather high axis ratio ($a/b > 1.6$) and long rotation period $\\sim 50$ hr in the dataset of 82P, though an interpretation in terms of such tidal effects is still speculative. \\subsection{Comets or asteroids?} At any rate, we have considered here only a small subset of the TSC events, focusing on the long TSCs/orbiters, for which there still exist only limited available data. Therefore we know little about such unique and extraordinary astronomical events, which are still full of ambiguities. In addition, we do not know whether QHCs are comets or asteroids. Indeed, several QHCs have sometimes been discovered or recovered as asteroids because of their cometary activity being weak: e.g., 36P (= 1925 QD = 1940 RP), 39P (= 1950 CR), 74P (= 1967 EU = 1978 NA$_6$ = 1981 UH$_{18}$ = 1982 YG$_3$), D/1977 C1 (= 1977 DV$_3$), P/1999 XN$_{120}$, P/2001 YX$_{127}$, and P/2003 CP$_7$. Further observations and research for the QHCs will be necessary to unlock their origin and nature." }, "0808/0808.2661_arXiv.txt": { "abstract": "{} {Using new Chandra X-ray observations and existing XMM-Newton X-ray and Hubble far ultraviolet observations, we aim to detect and identify the faint X-ray sources belonging to the Galactic globular cluster \\mbox{\\object{NGC 2808}} in order to understand their role in the evolution of globular clusters.} {We present a Chandra X-ray observation of the Galactic globular cluster \\mbox{\\object{NGC 2808}}. We classify the X-ray sources associated with the cluster by analysing their colours and variability. Previous observations with XMM-Newton and far ultraviolet observations with the Hubble Space Telescope are re-investigated to help identify the Chandra sources associated with the cluster. We compare our results to population synthesis models and observations of other Galactic globular clusters.} {We detect 113 sources, of which 16 fall inside the half-mass radius of \\mbox{\\object{NGC 2808}} and are concentrated towards the cluster core. From statistical analysis, these 16 sources are very likely to be linked to the cluster. We detect short-term (1~day) variability in X-rays for 7 sources, of which 2 fall inside the half-mass radius, and long-term (28~months) variability for 10 further sources, of which 2 fall inside the half-mass radius. Ultraviolet counterparts are found for 8 Chandra sources in the core, of which 2 have good matching probabilities and have ultraviolet properties expected for cataclysmic variables. We find one likely neutron star-quiescent low-mass X-ray binary and 7 cataclysmic variable candidates in the core of \\mbox{\\object{NGC 2808}}. The other 8 sources are cataclysmic variable candidates, but some could possibly be active binaries or millisecond pulsars. We find a possible deficit of X-ray sources compared to \\mbox{\\object{47 Tuc}} which could be related to the metallicity content and the complexity of the evolution of \\mbox{\\object{NGC 2808}}.} {} ", "introduction": "Globular clusters (GCs) are old, gravitationally bound stellar systems which can have extremely high stellar densities, especially in their core regions. In such an environment, dynamical interactions between the cluster members are inevitable, leading to a variety of close binary (CB) systems and other exotic stellar objects. The observed overabundance of neutron star (NS) low-mass X-ray binaries (LMXBs) in GCs relative to the Galactic field was explained by the dynamical processes occurring in the dense cores of GCs \\citep{fabian}. In contrast, evolution of a primordial binary into an LMXB in a GC is considered to be much less likely \\citep{VH87}. Observations also support the fact that quiescent LMXBs (qLMXBs) in GCs scale with the cluster encounter rate \\citep{GBW03,Pooley+03}, implying that qLMXBs are formed through dynamical processes in the dense cores. As white dwarfs (WDs) are far more common than NSs, we would then also expect many more close binaries containing an accreting WD primary, i.e. cataclysmic variables (CVs). The dynamically-formed CBs are expected to be found in the cores of GCs, where the stellar densities are at a maximum. The less dense regions outside the cores might be populated by CBs that evolved from primordial binaries \\citep[e.g.][]{Davies97} which are unlikely to survive in the dense core region. \\citet{HAS07} found that the combined effects of new binary creation and mass segregation exceed the destruction of primordial binaries in the central region of GCs, leading to a marked increase of the binary fraction in the central regions. Thus, we expect the majority of CBs, which are more massive than the mean stellar mass, to be located inside the half-mass radius. Outside the half-mass radius, the primordial binary fraction is well preserved \\citep{HAS07}. CBs are important for our understanding of GC evolution, since the binding energy of a few, very close binaries can rival that of a modest-sized GC \\citep[e.g.][and references therein]{EHI87,Hut+92,Hut+03}. In the core, binaries are subject to encounters and hard binaries become harder while transferring their energy to passing stars. Thus, CBs can significantly affect the dynamical evolution of the cluster. If there are only a few CBs, thermal processes dominate the cluster evolution, leading to core collapse followed by GC disruption on a timescale shorter than the mean age of GCs, estimated to be $12.9\\pm2.9$~Gyr \\citep{CGCFP00}. In contrast, the presence of many CBs leads to violent interactions, which heat the cluster, delay the core collapse, and promote its expansion. This depends critically on the number of CBs, which is still poorly known. Finding and studying these systems has proven to be extremely difficult, since the spatial resolution and detection limits of most available telescopes are insufficient for their detection. Only with the improved sensitivity and imaging quality of XMM-Newton and Chandra in the X-ray \\citep[e.g.][]{Webb+04,WWB06,Heinke+03,Heinke+06} and HST in the ultraviolet (UV) to infrared (IR) wavebands \\citep[ and references therein]{GHEM01,Albrow+01,EGHG03,EGHG03b,Knigge+02,Knigge+03} has it become possible to finally detect significant numbers of CB systems in GCs. The 13 bright X-ray sources found in the $\\sim$150 known Galactic GCs are LMXBs showing type I X-ray bursts \\mbox{\\citep[e.g.][]{1983adsx.conf...41L}}, whereas the faint sources belonging to the clusters are qLMXBs, CVs, active binaries (ABs, generally RS~CVn systems), or millisecond pulsars (MSPs). Multiwavelength studies can be used to identify the faint X-ray sources. For example, qLMXBs are usually identified by their soft blackbody-like or hydrogen atmosphere X-ray spectra \\mbox{\\citep[e.g.][]{GBW03,GBW03b}}, CVs can be confirmed by their blue, variable optical counterpart with hydrogen emission lines in their spectra \\mbox{\\citep[e.g.][]{Webb+04}}, ABs by their main-sequence, variable optical counterparts \\mbox{\\citep[e.g.][]{EGHG03}}, and MSPs by their radio counterpart \\mbox{\\citep[e.g.][]{GHEM01}}. Here, we present an X-ray study of the massive (\\mbox{$\\sim10^{6} M_{\\sun}$}) GC \\mbox{\\object{NGC 2808}} (\\mbox{$\\alpha = 09^{h} 12^{m} 02^{s}$, $\\delta = $-$64^{\\circ} 51{\\arcmin} 47{\\arcsec}$}). This intermediate metallicity GC \\citep[\\mbox{$\\rm{[Fe/H]} = -1.36$}, ][]{Walker99} lies at a distance of 9.6~kpc and is reddened by $E_{B-V}=0.22\\pm0.01$ \\citep{Harris96}. An absorption column of N${_H = 1.2\\times10^{21}\\mathrm{~cm^{-2}}}$ is derived from the reddening with the relation computed by \\citet{PS95}. The cluster has a very dense and compact core ($0.26\\arcmin$), a half-mass radius of $0.76\\arcmin$, a tidal radius of $15.55\\arcmin$, and a half-mass relaxation time of $1.35\\times10^{9}$~yrs \\citep{Harris96}. \\mbox{\\object{NGC 2808}} has received considerable attention in the literature and has been observed in the optical in detail as this GC is one of the most extreme examples with an unusual horizontal branch (HB) morphology, as first noted by \\citet{Harris74}. It shows a bimodal HB and one of the longest blue HB tails, the so-called extreme HB (EHB), with prominent gaps between the red HB (RHB), blue HB (BHB) and EHB \\citep[see also ][]{Bedin+00,Carretta+06}. Recently, \\citet{Piotto+07} found that \\mbox{\\object{NGC 2808}}'s main sequence (MS) is separated into three branches, which might be associated with the complex HB morphology and abundance distribution, and might be due to successive rounds of star formation with different helium abundances. \\mbox{\\object{NGC 2808}} is proposed as a good candidate to harbour an intermediate mass black hole (IMBH) in its core, due to its optical luminosity profile and EHB morphology \\citep{Miocchi07}. The core of \\mbox{\\object{NGC 2808}} has been imaged with the Space Telescope Imaging Spectrograph (STIS) on board the Hubble Space Telescope (HST) in the far-UV (FUV) and the near-UV (NUV). \\citet{Dieball+05} used the data set with an emphasis on the dynamically-formed stellar populations like CVs and blue stragglers (BSs) and young WDs. They found $\\sim$40~WD, $\\sim$60~BS and $\\sim$60~CV candidates in the field of view that covers the core of the cluster. Two of the CV candidates are variable (FUV sources~222 and 397). \\mbox{\\object{NGC 2808}} has also been observed with XMM-Newton in Feb. 2005. \\citet{Servillat+08-a} found 96~sources in the field of view (equivalent to the tidal radius), of which five fall inside the half-mass radius and are likely to be linked to the cluster. One qLMXB candidate and four CV candidates were discovered in the core of \\mbox{\\object{NGC 2808}}. However, several sources remained unresolved. In Sect.~\\ref{obs}, we present the new Chandra X-ray data, and then compare them to the XMM-Newton observations (Sect.~\\ref{xmm}). We present HST and XMM-Newton Optical Monitor UV counterparts in Sect.~\\ref{uv}. We finally discuss our results in Sect.~\\ref{disc}. ", "conclusions": "We presented Chandra observations of \\mbox{\\object{NGC 2808}} coupled with previous XMM-Newton observations \\citep{Servillat+08-a}, HST FUV observations \\citep{Dieball+05}, VLT and ATCA observations. We have shown that 16 Chandra sources are likely to be linked to the cluster, with possibly a 17th close to the half mass radius. One of these is consistent with the X-ray emission of a qLMXB, confirming the previous detection with XMM-Newton. Two Chandra sources (7 and 14) have FUV counterparts that show emission compatible with a CV. Chandra source 10 is likely to be a CV from its X-ray emission, but no UV counterparts was found to confirm its nature. Another highly variable source (16) in the core is likely to be a CV, as well as two other variable sources (Chandra 3 and XMM-Newton C5). Two other Chandra sources (8 and 11) have optical counterparts compatible with the expected emission of CVs. We have thus identified 7 CV candidates (plus XMM-Newton source C5) and the observations indicate that there may be as many as 15 in the Chandra observations (although some of the faintest may be ABs or MSPs), along with $\\sim30$ CV candidates in the HST UV observations. This significant population of close binaries is likely to play an important role in slowing down the core collapse of this cluster. Compared to the number of X-ray sources detected in \\mbox{\\object{47 Tuc}} and expected from dynamical formation, we found a possible deficit of X-ray sources in \\mbox{\\object{NGC 2808}}. This might indicate a true deficit of CVs, which is possibly linked to the metallicity content and the complexity of the evolution of \\mbox{\\object{NGC 2808}}." }, "0808/0808.1512_arXiv.txt": { "abstract": "We examine the temperature-dependent electroweak phase transition in extensions of the Standard Model in which the electroweak symmetry is spontaneously broken via strongly coupled, nearly-conformal dynamics. In particular, we focus on the low energy effective theory used to describe Minimal Walking Technicolor at the phase transition. Using the one-loop effective potential with ring improvement, we identify significant regions of parameter space which yield a sufficiently strong first order transition for electroweak baryogenesis. The composite particle spectrum corresponding to these regions can be produced and studied at the Large Hadron Collider experiment. We note the possible emergence of a second phase transition at lower temperatures. This occurs when the underlying technicolor theory possesses a nontrivial center symmetry. ", "introduction": "The experimentally observed baryon asymmetry of the universe may be generated at the electroweak phase transition (EWPT) \\cite{Shaposhnikov:1986jp,Shaposhnikov:1987tw,Shaposhnikov:1987pf, Farrar:1993sp,Farrar:1993hn,Gavela:1993ts,Gavela:1994ds}. {}For the mechanism to be applicable it requires the presence of new physics beyond the Standard Model (SM) \\cite{Nelson:1991ab,Joyce:1994bi,Joyce:1994fu,Joyce:1994zn, Joyce:1994zt,Cline:1995dg}. {An essential condition for electroweak baryogenesis is that the baryon-violating interactions induced by electroweak sphalerons are sufficiently slow immediately after the phase transition to avoid the destruction of the baryons that have just been created. This is achieved when the thermal average of the Higgs field evaluated on the ground state, in the broken phase of the electroweak symmetry, is large enough compared to the critical temperature at the time of the transition (see for example ref.~\\cite{Cline:2006ts} and references therein), \\beq \\phi_c/ T_c > 1. \\label{cond} \\eeq In the SM, the bound (\\ref{cond}) was believed to be satisfied only for very light Higgs bosons \\cite{Carrington:1991hz,Arnold:1992fb,Arnold:1992rz, Anderson:1991zb,Dine:1992wr}. However, this was before the mass of the top quark was known. With $m_t=175$ GeV, nonperturbative studies of the phase transition \\cite{Kajantie:1995kf} show that the bound (\\ref{cond}) cannot be satisfied for {\\it any} value of the Higgs mass. In addition to the difficulties with producing a large enough initial baryon asymmetry, the impossibility of satisfying the sphaleron constraint (\\ref{cond}) in the SM (Standard Model) provides an incentive for seeing whether the situation improves in various extensions of the SM. We refer to \\cite{Cline:2006ts} for a summary of the different attempts in this direction. In this paper we explore the electroweak phase transition in a model in which the electroweak symmetry is broken dynamically \\cite{Weinberg:1979bn,Susskind:1978ms}. A dynamical origin behind the spontaneous breaking of the electroweak symmetry is a natural extension of the SM. However, electroweak precision data and constraints from flavor changing neutral currents both disfavor an underlying gauge dynamics resembling too closely a scaled-up version of Quantum Chromodynamics (QCD) (see \\cite{Sannino:2008ha,Hill:2002ap,Lane:2002wv} for recent reviews). Since technicolor models have been less fashionable than supersymmetric models in the last decade, it is worthwhile to review the recent progress that has enhanced their attractiveness from the particle physics perspective. One area of progress is in the understanding of the phase diagram \\cite{Sannino:2004qp,Dietrich:2006cm,Ryttov:2007sr,Ryttov:2007cx}, as function of the number of flavors and colors, of any SU(N) non-supersymmetric gauge theory with fermionic matter transforming according to various representations of the underlying gauge group. This has made it possible to provide the first classification of the possible theories one can use to break the electroweak symmetry \\cite{Dietrich:2005jn,Dietrich:2006cm}. New analytic tools such as the all-order beta function \\cite{Ryttov:2007cx} allow the determination, for the first time, of the anomalous dimension of the mass of the fermions at the nonperturtative infrared fixed point. This information is crucial for walking technicolor models \\cite{Holdom:1984sk,Eichten:1979ah,Holdom:1981rm,Yamawaki:1985zg,Appelquist:1986an,Lane:1989ej}, {\\it i.e.}, the ones for which the underlying gauge dynamics is nearly conformal. A key realization that enabled further progress was that gauge theories with fermions in two-index (symmetric or adjoint) representations of the underlying gauge group have interesting features \\cite{Sannino:2004qp,Dietrich:2005jn,Dietrich:2006cm,Ryttov:2007sr,Ryttov:2007cx}, such as the possibility of the existence of a nonperturbative infrared fixed point for a very low number of flavors \\cite{Sannino:2004qp}, naturally reducing the tension with precision data \\cite{Sannino:2004qp,Dietrich:2005jn,Foadi:2007ue,Foadi:2007se}. These properties make them intriguing candidates for walking technicolor type models \\cite{Sannino:2004qp,Dietrich:2005jn} (related studies can be found in \\cite{Christensen:2005cb}). In contrast, the naive scaling up of QCD, which is far from conformal, is strongly contradicted by phenomenological constraints \\footnote{The reader will find in Appendix F of \\cite{Sannino:2008ha} a complete account of alternative large N limits one can use to gain information on the spectrum of theories with matter in higher dimensional representation}. Another important development occurred in first principle lattice simulations of the minimal walking technicolor theories, carried out in refs.\\ \\cite{Catterall:2008qk, Catterall:2007yx,Shamir:2008pb,DelDebbio:2008zf,DelDebbio:2008wb}. These studies give preliminary support to the analytical arguments that these theories are near or actually already conformal. The case of fermions in the fundamental representation has been investigated in \\cite{Catterall:2008qk,Appelquist:2007hu,Deuzeman:2008sc}. On the astrophysical side, technicolor models are capable of providing interesting dark matter candidates, since the new strong interactions confine techniquarks in technimeson and technibaryon bound states. The spin of the technibaryons depends on the representation according to which the technifermions transform, and the numbers of flavors and colors. The lightest technimeson is short-lived, thus evading BBN constraints, but the lightest technibaryon has typically \\footnote{there may be situations in which the technibaryon is a goldstone boson of an enhanced flavor symmetry} a mass of the order \\begin{equation} m_{TB} \\sim 1-2\\ {\\rm TeV} \\ . \\end{equation} Technibaryons are therefore natural dark matter candidates \\cite{Nussinov:1985xr,Barr:1990ca,Gudnason:2006yj}. In fact it is possible to {naturally} understand the observed ratio of the dark to luminous matter mass fraction of the universe if the technibaryon possesses an asymmetry \\cite{Nussinov:1985xr,Barr:1990ca,Gudnason:2006yj}. If the latter is due to a net $B-L$ generated at some high energy scale, then this would be subsequently distributed among {\\em all} electroweak doublets by fermion-number violating processes in the SM at temperatures above the electroweak scale \\cite{Shaposhnikov:1991cu,Kuzmin:1991ft,Shaposhnikov:1991wi}, thus naturally generating a technibaryon asymmetry as well. To avoid experimental constraints the technibaryon should be constructed in such a way as to be a complete singlet under the electroweak interactions \\cite{Barr:1990ca,Dietrich:2006cm} while still having a nearly conformal underlying gauge theory \\cite{Dietrich:2006cm}. In this case it would be hard to detect it in current earth-based experiments such as CDMS \\cite{Bagnasco:1993st,Gudnason:2006yj,Kouvaris:2008hc,Akerib:2004fq,Akerib:2005kh}. Other possibilities have been envisioned in \\cite{Kouvaris:2007iq,Khlopov:2007ic} and possible astrophysical effects studied in \\cite{Kouvaris:2007ay}. One can alternatively obtain dark matter from possible associated new sectors instead of the technicolor sector \\cite{Kainulainen:2006wq}, including those which are not gauged under the electroweak interactions \\cite{Dietrich:2006cm}. In \\cite{Sannino:2008ha} the reader will find an up-to-date summary of the recent efforts in this direction. Coming to the main topic of this paper, the order of the electroweak phase transition (EWPT) depends on the underlying type of strong dynamics and plays an important role for baryogenesis \\cite{Cline:2002aa,Cline:2006ts}. The technicolor chiral phase transition at finite temperature is mapped onto the electroweak one. Attention must be paid to the way in which the electroweak symmetry is embedded into the global symmetries of the underlying technicolor theory. An interesting preliminary analysis dedicated to earlier models of technicolor has been performed in \\cite{Kikukawa:2007zk}. In this work, we wish to investigate the EWPT in a class of realistic and viable technicolor models. An explicit phenomenological realization of walking models consistent with the electroweak precision data is termed Minimal Walking Technicolor (MWT) \\cite{Foadi:2007ue}. It is based on an SU(2) gauge theory coupled to two flavors of adjoint techniquarks. This model is thought to lie close, in theory space, to theories with nontrivial infrared fixed points \\cite{Sannino:2004qp,Ryttov:2007cx}. Indeed it is possible that it already has such a fixed point itself. In the vicinity of such a zero of the beta function, the coupling constant flows slowly (``walks''). This theory possesses an SU(4) global symmetry. At the LHC one will observe the composite states which are classified according to irreducible representations of the stability group left invariant by the technifermion condensate. The stability group, here, corresponds to the SO(4) symmetry which contains the SU(2) custodial symmetry of the SM. We choose the natural SM embedding, as detailed in the following section. In ref.\\ \\cite{Foadi:2007ue} a comprehensive Lagrangian was introduced for this model, taking into account the global symmetries of the underlying gauge theory, the walking dynamics via the modified Weinberg sum rules \\cite{Appelquist:1998xf}, and the constraints coming from precision data \\cite{Foadi:2007se}. The effective theory contains composite scalars and spin-one vectors. Compatibility between the electroweak precision constraints and tree-level unitarity of $WW$-scattering was demonstrated in \\cite{Foadi:2008ci}. The study of longitudinal WW scattering unitarity versus precision measurements within the effective Lagrangian approach demonstrated that it is possible to pass the precision tests while simultaneously delaying the onset of unitarity \\cite{Foadi:2008ci}. In the present work we will use as a template the low energy effective theory developed in \\cite{Foadi:2007ue}. We start in section \\ref{sect2} by summarizing the basic theory, highlighting the degrees of freedom relevant near the phase transition. In section \\ref{sect3} the finite-temperature effective potential is then computed at the one-loop order, including the resummation of ring diagrams. Our analysis is presented in section \\ref{sect4}. As a preliminary investigation we adopt the high-temperature expansion results for the effective potential. We then explore the region of the effective theory parameters yielding a first-order phase transition and study its strength. The ratio of the composite Higgs thermal expectation value at the critical temperature divided by the corresponding temperature is determined as function of the parameters of the low energy effective theory. We identify a significant region of parameter space where this ratio is sufficiently large to induce electroweak baryogenesis. {The spectrum of the composite spin-zero states directly associated to these regions can be investigated and the related particles produced at the Large Hadron Collider experiment. In the subsection \\ref{sect4.d} we note the possible emergence of a second phase transition at lower temperatures, {\\it i.e.}, the confinement/deconfinement one. This transition occurs when the underlying technicolor theory possesses a nontrivial center symmetry.} Several appendices are provided, which give details concerning our analytical results. ", "conclusions": "" }, "0808/0808.3384_arXiv.txt": { "abstract": "Despite compelling arguments that significant discoveries of physics beyond the standard model are likely to be made at the Large Hadron Collider, it remains possible that this machine will make no such discoveries, or will make no discoveries directly relevant to the dark matter problem. In this article, we study the ability of astrophysical experiments to deduce the nature of dark matter in such a scenario. In most dark matter studies, the relic abundance and detection prospects are evaluated within the context of some specific particle physics model or models ({\\textit e.g.,} supersymmetry). Here, assuming a single weakly interacting massive particle constitutes the universe's dark matter, we attempt to develop a model-independent approach toward the phenomenology of such particles in the absence of any discoveries at the Large Hadron Collider. In particular, we consider generic fermionic or scalar dark matter particles with a variety of interaction forms, and calculate the corresponding constraints from and sensitivity of direct and indirect detection experiments. The results may provide some guidance in disentangling information from future direct and indirect detection experiments. ", "introduction": "The consensus of the astrophysics community is that a large fraction of the universe's mass consists of non-luminous, non-baryonic material, known as dark matter \\cite{Bertone:2004pz}. Although the nature of this substance or substances remains unknown, weakly interacting massive particles (WIMPs) represent a particularly attractive and well motivated class of possibilities. Although the most studied WIMP candidate is the lightest neutralino \\cite{neutralino} in supersymmetric models, many other possibilities have also been proposed, including Kaluza-Klein states in models with universal \\cite{universal,universal2} or warped \\cite{warped} extra dimensions, stable states in Little Higgs theories \\cite{tparity}, and many others. In each of the above mentioned cases, many new particle species, in addition to the WIMP itself, are expected to lie within the discovery reach of the Large Hadron Collider (LHC), making the task of deducing the nature of the WIMP immeasurably simpler. In supersymmetry, for example, gluinos and squarks are expected to be produced prolifically. By studying the cascades produced in the decays of such particles, the masses of several superparticle masses, including the lightest neutralino, are likely to be determined. If squarks, gluinos, and other additional superpartners are too heavy to be produced, however, the lightest neutralino will also be very difficult to study at the LHC, even if rather light itself. More generally speaking, in the absence of heavier particles with shared quantum numbers, WIMPs will not be easily detected or studied at the LHC. Although an electroweak scale, cold thermal relic particle, if it exists, would almost certainly be produced at the LHC, identifying and characterizing the nature of the WIMP simply from missing energy studies is a daunting, perhaps impossible, task \\cite{Birkedal:2004xn,Feng:2005gj}. Although the usual list of prospective WIMPs mentioned above contains some very attractive and well-motivated candidates for dark matter, there are certainly many possible forms of dark matter that have not yet been considered. As the first observations of particle dark matter might well come from direct and/or indirect detection experiments, it is possible that these results may be misinterpreted as a result of theoretical bias, anticipating dark matter to have the properties of a neutralino or other often studied candidates. To avoid such confusion, model-independent studies of dark matter phenomenology can play an important role (for previous work in this direction, see Refs.\\ \\cite{Birkedal:2004xn,Kurylov:2003ra,Giuliani:2004uk}). In this article, rather than consider a WIMP candidate from a specific theoretical model, we study model-independent WIMPs with different combinations of spins and interaction forms with standard model particles. These interactions are limited only by the requirements of Lorentz invariance and a consequent WIMP abundance consistent with cosmological observations. For each spin and interaction form, we evaluate the constraints from and prospects for direct and indirect detection of WIMPs in current and future experiments. Although we will be forced to adopt some assumptions in order to make the problems at hand tractable, we attempt to be as general as possible throughout our study. Beyond the starting point that the dark matter is a WIMP in the form of a single species of a cold thermal relic, we adopt only two assumptions: \\begin{enumerate} \\item Any new particle species in addition to the WIMP has a mass much larger than the WIMP. \\item The WIMP interactions with standard model particles are dominated by those of one form (scalar, vector, \\textit{etc.}). \\end{enumerate} An implication of the first assumption is that the WIMP's thermal abundance is not affected by resonances or coannihilations. At a later stage of this paper, we will discuss the impact of relaxing these assumptions. The remainder of this paper is structured as follows. In Sec.\\ \\ref{fermion} we explore the phenomenology of a generic fermionic WIMP, including its annihilation cross section and relic abundance, elastic scattering cross section and direct detection prospects, and indirect detection prospects in the form of a neutrino flux from the Sun and gamma rays and charged particles produced in galactic annihilations. In Sec.\\ \\ref{scalar}, we repeat this exercise for the case of a scalar WIMP. In each of these two sections, we also consider dark matter candidates from specific particle physics frameworks and discuss how they fit into our model-independent analysis. In Sec.\\ \\ref{conclusion} we summarize our results and present our conclusions. ", "conclusions": "Even if the Large Hadron Collider does not reveal physics beyond the standard model, a dark matter candidate in the form of a weakly interacting massive particle may still exist. In this article, we have studied how the nature of such a WIMP could be deduced by its signatures in astrophysical experiments. In our analysis, we have taken a general and model-independent approach, considering fermionic or scalar WIMPs with a variety of interaction forms. In Table \\ref{t1}, we summarize our findings. For each combination of spin and interaction form, we indicate the constraints placed by and the prospects for direct detection experiments, high-energy neutrino telescopes, and indirect detection experiments using gamma-rays or charged particles. Under the column of direct detection, we use the phrases ``strongly excluded,'' ``weakly excluded,'' or ``within near future reach,'' to denote the sensitivity or prospects for each case. By ``strongly excluded,'' we indicate instances in which the effective couplings to quarks, as relevant to elastic scattering with nuclei, must be suppressed by more than a factor of ten relative to the value required to generate thermally the observed dark matter abundance. As we have discussed, such a suppression could result from resonant annihilations, coannihilations, or annihilations to gauge/Higgs bosons. The label ``weakly excluded,'' in contrast, indicates only that the case is excluded if the effective couplings to quarks are not suppressed by such effects. Lastly, the label ``within near-future reach'' indicates an elastic scattering cross section (without suppression) that is within approximately two order of magnitudes of current direct detection limits. Under the column of neutrino telescopes, we classify each case as either not sensitive or sensitive over a range of WIMP masses (for next generation experiments, such as IceCube). This evaluation depends on the annihilation products of the WIMP, however, and thus are highly approximate. Under the column of gamma-rays and charged cosmic rays, we simply indicate whether the WIMP's annihilations are or are not suppressed by the square of the WIMP's velocity. If such velocity suppression is present, it is highly unlikely that GLAST, PAMELA or other planned indirect detection experiments will be capable of detecting dark matter. This leads us to the most obvious and important question: Will the information provided by direct and indirect detection experiments be able to be used to infer the particle nature of the dark matter? Although there are certainly cases in which measurements by these experiments will not lead to an unambiguous identification, there are many scenarios in which a great deal could be learned. For example, if IceCube or another high-energy neutrino telescope were to observe neutrinos from WIMP annihilations in the Sun, we would be able to conclude that the WIMP is likely fermionic,\\footnote{More precisely, we could conclude in this case that the dark matter particle is not a scalar. Vector WIMPs, which we have not studied in this paper, could also potentially generate an observable flux of high-energy neutrinos \\cite{Hooper:2002gs}.} and that it possesses an axial interaction with light quarks. By studying the precise rate observed, one could also potentially determine whether the WIMP's axial interaction played a dominant or only subdominant role in the process of thermal freeze-out. This could be combined with observations from direct detection experiments to further constrain the possible interactions possessed by the WIMP. As a second possible scenario, imagine that near future direct detection experiments observe a WIMP with a mass of a few hundred GeV and that, shortly afterward, GLAST observes a corresponding gamma-ray signal from WIMPs annihilating in the halo. From Table \\ref{t1}, we can see that this leads us to only three likely possibilities: the WIMP is either a fermion with vector interactions, a fermion with pseudoscalar-scalar interactions, or a scalar with Yukawa-like scalar interactions. Although previous studies have shown that dark-matter experiments have the potential to constrain the parameters of supersymmetry \\cite{susydetermine} or even to help identify the theoretical framework from which the dark matter arises \\cite{otherdetermine}, here we have demonstrated that a far more model-independent approach can also be fruitful. In particular, without assuming any particular theoretical framework or model, we have shown that direct and indirect dark matter experiments can be used to considerably constrain the spin and interactions of the dark matter, even in the absence of any discoveries at the LHC. \\vspace{0.5cm} \\begin{table*}[!ht] \\hspace{0.0cm} \\begin{center} \\begin{large} Fermionic Dark Matter \\end{large} \\end{center} \\begin{ruledtabular} \\begin{tabular} {c c c c c c c c} Interaction &\\vline& Direct Detection & \\,\\vline \\, & Neutrino Telescopes &\\,\\vline \\, & $\\gamma$-rays, $e^{\\pm}$, $\\bar{p}$ & \\\\ \\hline \\hline Scalar &\\vline& Strongly Excluded $M_{\\chi} \\approx 10-100$ GeV & \\vline & Not Sensitive & \\vline & Suppressed by $v^2$ & \\\\ ($G_f \\propto m_f$) &\\vline& Weakly Excluded $M_{\\chi} \\approx 100-200$ GeV &\\vline & & \\vline & & \\\\ &\\vline& Within Near Future Reach $M_{\\chi} \\approx 200-300$ GeV &\\vline & & \\vline & & \\\\ \\hline Scalar &\\vline& Strongly Excluded $M_{\\chi} \\approx 10$ GeV$-10$ TeV & \\vline & NA & \\vline& Suppressed by $v^2$ & \\\\ ($G_f$ {\\rm Universal}) &\\vline& & \\vline & & \\vline& & \\\\ \\hline Pseudoscalar &\\vline& Not Sensitive & \\vline & Not Sensitive & \\vline& Unsuppressed & \\\\ \\hline Vector/Tensor &\\vline& Strongly Excluded $M_{\\chi}\\approx 10-350$ GeV & \\vline & Not Sensitive & \\vline& Unsuppressed & \\\\ &\\vline& Weakly Excluded $M_{\\chi} \\approx 350$ GeV$-2$ TeV & \\vline & & \\vline & & \\\\ \\hline Axial &\\vline& Not Sensitive & \\vline & Sensitive $M_{\\chi} \\sim 100-500$ GeV & \\vline& Suppressed by $v^2$ & \\\\ \\hline Scalar-Pseudoscalar &\\vline& Not Sensitive & \\vline & Not Sensitive & \\vline& Suppressed by $v^2$ & \\\\ \\hline Pseudoscalar-Scalar &\\vline& Weakly Excluded $M_{\\chi} \\approx 10-180$ GeV & \\vline & Not Sensitive & \\vline& Unsuppressed & \\\\ ($G_f \\propto m_f$) &\\vline& Within Near Future Reach $M_{\\chi} \\approx 180-800$ GeV &\\vline & & \\vline & & \\\\ \\hline Vector-Axial &\\vline& Not Sensitive & \\vline & Not Sensitive & \\vline& Unsuppressed & \\\\ \\hline Axial-Vector &\\vline& Strongly Excluded $M_{\\chi} \\approx 10$ GeV$-2$ TeV & \\vline & Not Sensitive & \\vline& Unsuppressed & \\\\ &\\vline& Weakly Excluded $M_{\\chi} \\approx 2-10$ TeV & \\vline & & \\vline & & \\\\ \\end{tabular} \\end{ruledtabular} \\vspace{0.5cm} \\begin{center} \\begin{large} Scalar Dark Matter \\end{large} \\end{center} \\begin{ruledtabular} \\begin{tabular} {c c c c c c c c} Interaction &\\vline& Direct Detection & \\,\\vline \\, & Neutrino Telescopes &\\,\\vline \\, & $\\gamma$-rays, $e^{\\pm}$, $\\bar{p}$ & \\\\ \\hline \\hline Scalar &\\vline& Weakly Excluded $M_{\\phi} \\approx 10-70$ GeV & \\vline & Not Sensitive & \\vline & Unsuppressed & \\\\ ($F_f \\propto m_f$) &\\vline& Within Near Future Reach $M_{\\phi} \\approx 70-200$ GeV &\\vline & & \\vline & & \\\\ \\hline Scalar &\\vline& Strongly Excluded $M_{\\phi} \\approx 10$ GeV$-10$ TeV & \\vline & NA & \\vline& Unsuppressed & \\\\ ($F_f$ {\\rm Universal}) &\\vline& & \\vline & & \\vline& & \\\\ \\hline Vector &\\vline& Strongly Excluded $M_{\\phi}\\approx 10$ GeV$-1$ TeV & \\vline & Not Sensitive & \\vline& Suppressed by $v^2$ & \\\\ &\\vline& Weakly Excluded $M_{\\phi} \\approx 1-5$ TeV &\\vline & & \\vline & & \\\\ \\hline Scalar-Pseudoscalar &\\vline& Not Sensitive & \\vline & Not Sensitive & \\vline & Unsuppressed & \\\\ \\hline Vector-Axial &\\vline& Not Sensitive & \\vline & Not Sensitive & \\vline & Suppressed by $v^2$ & \\\\ \\end{tabular} \\end{ruledtabular} \\caption{A summary of our results, describing the sensitivity and prospects for the direct and indirect detection of dark matter particles in the various cases we have considered. See the text for explanations for the labels used.} \\label{t1} \\end{table*} The results presented in Table \\ref{t1} rely upon the set of assumptions we have adopted. It must be noted that if dark matter consists of non-thermally produced WIMPs, or of multiple species of particles, our conclusions could be altered considerably. Furthermore, one might worry that the effects of resonances, coannihilations, or annihilations to gauge/Higgs bosons, which we have largely neglected in our analysis, might dramatically change our conclusions. To some degree, however, the impact of such processes are encapsulated in our definitions of ``strongly excluded'' and ``weakly excluded'', as used in Table \\ref{t1}. For example, if a WIMP annihilates largely through a narrow resonance such that twice the mass of the WIMP lies within approximately 5\\% of the exchanged particle, then the effective couplings relevant for elastic scattering can be reduced by a factor of ten without the WIMP being overproduced in the early universe (see Sec.\\ \\ref{resonance}). This mildly (5\\% or less) fine-tuned resonance corresponds to the ``weakly excluded'' label used in the table. Anything labeled ``strongly excluded'' would require the masses to be tuned even more precisely to the resonance value to remain viable. Similarly, if a significant fraction of WIMP annihilations in the early universe proceeded to a combination of gauge or Higgs bosons, or occurred through coannihilations with another species of particle, the elastic scattering cross section could be suppressed. For the scenarios we have labeled as ``strongly excluded'' to have remained hidden from direct detection experiments, however, about $99$\\% or more of the annihilations/coannihilations of WIMPs in the early universe must have occurred through such processes. So although the conclusions we have reached here are not entirely immune to the inclusion of such effects, they are quite robust in all but the most extreme cases. In conclusion, we find that in the case that the Large Hadron Collider does not discover physics beyond the standard model, astrophysical experiments may still be able to constrain the nature of the dark matter, even without assuming supersymmetry or any other specific particle physics framework. In particular, the spin and interaction forms of dark matter can potentially be identified by combining results from direct detection experiments, neutrino telescopes, and indirect detection experiments using gamma-rays or charged cosmic rays. \\vspace{0.5cm}" }, "0808/0808.3451_arXiv.txt": { "abstract": "% {}{While observational evidence shows that most of the decline in a star's X-ray activity occurs between the age of the Hyades ($\\sim 8 \\times 10^8$ yrs) and that of the Sun, very little is known about the evolution of stellar activity between these ages. To gain information on the typical level of coronal activity at a star's intermediate age, we studied the X-ray emission from stars in the 1.9 Gyr old open cluster NGC~752.} {We analysed a $\\sim$ 140 ks \\chandra\\ observation of NGC~752 and a $\\sim$ 50 ks \\xmm\\ observation of the same cluster. We detected 262 X-ray sources in the \\chandra\\ data and 145 sources in the \\xmm\\ observation. Around 90\\% of the catalogued cluster members within \\chandra's field-of-view are detected in the X-ray. The X-ray luminosity of all observed cluster members (28 stars) and of 11 cluster member candidates was derived.} {Our data indicate that, at an age of 1.9 Gyr, the typical X-ray luminosity of the cluster members with $M=0.8-1.2~M_{\\sun}$ is $L_{\\rm X} = 1.3 \\times 10^{28}$ \\es, so approximately a factor of 6 less intense than that observed in the younger Hyades. Given that $L_{\\rm X}$ is proportional to the square of a star's rotational rate, the median $L_{\\rm X}$ of NGC~752 is consistent, for $t \\ga 1$ Gyr, to a decaying rate in rotational velocities $v_{\\rm rot} \\propto t^{-\\alpha}$ with $\\alpha \\sim 0.75$, steeper than the Skumanich relation ($\\alpha \\simeq 0.5$) and significantly steeper than observed between the Pleiades and the Hyades (where $\\alpha <0.3$), suggesting that a change in the rotational regimes of the stellar interiors is taking place at $t\\sim 1$ Gyr.}{} ", "introduction": "\\label{sec:intro} Investigations of solar-type stars in the better-studied nearby open clusters have provided a basis for understanding the evolution of stellar activity and rotation velocity. The early {\\it Einstein} and {\\sc Rosat} studies of open clusters revealed that the mean X-ray luminosity of G, K and early M stars decreases steadily with age from pre-main sequence stars (PMS) through the Pleiades ($\\sim 8\\times 10^7$ yrs), the Hyades ($\\sim 8\\times 10^8$ yrs) and the old disc population (\\citealp{randich2000}; \\citealp{micela2002}). In main-sequence stars, late-type stellar magnetic activity is regulated principally by rotation (\\citealp{pgr+81}; \\citealp{bds+95}) and its decay with stellar age is attributed to rotational spin-down. Magnetic braking models indicate that surface rotational velocities should decline as $v_{\\rm rot} \\propto t^{-\\alpha}$ with $0.38 < \\alpha < 0.75$, depending on whether the magnetic field geometry is radial or closer to a dipole configuration (\\citealp{kawaler88}). Models considering possible differences in the pre-main sequence disc-locking time, solid-body versus differential rotation in the interior, and the onset of magnetic saturation give decay laws over the range $0.2 < \\alpha< 0.8$ for 1$-$5 Gyr old solar-mass stars (\\citealp{kpb+97}). Although \\cite{skuma72} found the decay in rotational velocities and the strength of the magnetic activity indicators Ca\\,{\\sc ii} to be consistent with pure magnetic breaking ($\\alpha \\simeq 0.5$) from the age of the Pleiades to that of the Sun, more recent evidence points to a decay rate of the stellar activity which is less steep between the Pleiades and the Hyades and drops significantly faster from the Hyades ages to that of the Sun or field stars (e.g. \\citealp{shb85} for the evolution of transition-region lines; \\citealp{micela02} and \\citealp{pf05} for the evolution of coronal emission). This may indicate a change, between 1 and 5 Gyr, in the rotational regime of stellar interiors. Helioseismology has shown that in the Sun there is no difference between the angular rotation of the radiative core and that of the convective envelope, while models suggest that a significant radial gradient is present at earlier stages (\\citealp{mb91}, \\citealp{bcm99}, \\citealp{silpin00}). \\cite{qam+98} reported the surface rotational evolution between the Pleiades and the Hyades to be less steep than the \\cite{skuma72} relation, and interpreted this as evidence of angular momentum transport from the core to the envelope. Determination of the evolution of coronal emission beyond the Hyades age can therefore be an effective probe of the evolution of stellar interiors in older stars. Although most of the decline in X-ray activity occurs between the age of the Hyades and that of the Sun, the evolution of stellar activity in solar-type stars older than the Hyades is still poorly known. For stars older than the Hyades, the data on coronal emission level are based on field stars and the Sun. While the Sun's age is well constrained at 4.5 Gyr, it is only one star and therefore cannot be assumed to be representative of the behavior of solar-type stars at its current age. The age of field stars are poorly constrained and they represent in general a mixture of stellar populations and chemical composition. The best approach to explore the stellar activity of solar-type stars at an age intermediate between those of the Hyades and the Sun is the X-ray observation of an old open cluster, which naturally provides a sample of similar aged stars, at the same distance and with similar chemical abundances. This obvious approach is difficult for observational reasons: there are few old clusters in the solar neighbourhood, since many of them are dissipated in few billion years through dynamical evolution, and their intrinsically low X-ray luminosity requires extremely long exposures with current X-ray instrumentation. The only known, sufficiently nearby, intermediate age, open cluster is NGC~752. At a distance of $\\sim$ 400 pc and with age between 1.7 and 2 Gyr, this is the only open cluster in which one can measure the X-ray luminosity of individual dG stars with exposure times below 200 ks -- unfortunately, though, the cluster is rather dispersed. We present here the results of the analysis of two X-ray observations of NGC~752: a $\\sim$ 140 ks \\chandra\\ observation and a $\\sim$ 50 ks \\xmm\\ observation. The analysis focuses on solar-type stars, with mass $0.8 < M < 1.2~M_{\\sun}$. The paper is organised as follows. After a summary, below, of the properties of NGC~752, the observations and data analysis are described in Sect.\\,\\ref{sec:obs}. Results are presented in Sect.\\,\\ref{sec:res} and the implications are discussed in Sect.\\,\\ref{sec:disc}. \\subsection{NGC~752} \\label{sec:ngc752} The most comprehensive study of the open cluster NGC~752 to date is the one by \\cite{dlm+94} (hereafter DLM94). These authors collected and systematised a number of existing proper-motion and radial-velocity studies of this cluster, supplemented them with their own measures of radial velocities, and combined them with existing photometric and spectroscopic data. Thanks to this effort, they claim to have obtained an effectively complete census of all probable and possible cluster members down to the observable limit of the unevolved main sequence which, in their case, corresponds to $V \\sim 13.5-14$. The electronic version of their catalogue available from {\\sc Vizier} lists 255 stars of which 157 are classified as probable or possible members of NGC~752. Combining the spectroscopic and photometric approaches, DLM94 establish the cluster metallicity to be [Fe/H] $= -0.15 \\pm 0.05$ dex. They also derive a reddening value of $E(B-V) = 0.035 \\pm 0.005$, which corresponds to an absorbing column density of $N({\\rm H}) = 2 \\times 10^{20}$ \\cmtwo\\ or $A_V = 0.1$ (\\citealp{cox00}). From intermediate-band photometry and main-sequence fitting, DLM94 derive a distance modulus of $(m-M)=8.25 \\pm 0.10$ mag which corresponds to a distance of $\\sim 430$ pc (assuming $A_V = 0.1$). Comparing the data with theoretical models, they estimate for this cluster an age of $1.9\\pm0.2$ Gyr from classical isochrones and $1.7\\pm0.1$ from isochrones with over-shooting, in good agreement with the results from isochrone fitting of \\cite{mmm93} (1.8 Gyr) and of \\cite{ddg+95} (2.0 Gyr). Since the first proper-motion study by \\cite{ebbi39}, it has been known that in NGC~752 the unevolved main sequence is deficient in stars relative to the turn-off. This is likely due to mass-segregation effects and the cluster relaxation, which leads to the evaporation of the less massive members. NGC~752 was observed in the X-ray with {\\sc Rosat} by \\cite{bv96}. They catalogued 49 sources, seven of which were identified with cluster members. ", "conclusions": "\\label{sec:disc} The aim of this paper is to determine the level of stellar activity of solar type stars at an intermediate age between that of the Hyades and those of field stars. In Fig.\\,\\ref{fig:lxvage}, we compare the luminosities of members of NGC~752 with masses in the range $0.8-1.2~M_{\\sun}$ with the luminosity distributions of the Pleiades, the Hyades and the field stars. These three points are from \\cite{micela02} and refer to stars within the same mass range ($0.8-1.2~M_{\\sun}$). As summarised in Table\\,\\ref{tab:lx}, in NGC~752 we detected in the X-ray 6 cluster members within this mass range (the points in the figure), 5 stars within the \\chandra\\ FOV and one outside (significantly brighter than the others) detected by \\xmm. For stars detected both by \\chandra\\ and \\xmm\\ we used in the figure the \\chandra\\ derived flux estimate. Within the \\chandra\\ FOV there is one additional known member with solar mass: Id 790 in the work by DLM94 (easily identifiable in the CMD in Fig.\\,\\ref{fig:colmag}). The upper limit on the X-ray luminosity of this star is also given in the figure. In the figure, the crosses give the median value of the distributions (including upper limits) and the size of the vertical bars are determined by the 25\\% and 75\\% quantiles. In Sect.\\,\\ref{sec:res} we showed that we expect the optical sample by DLM94 to be complete within the \\chandra\\ FOV for $M > 0.8~M_{\\sun}$, thus the sample of the 5 \\chandra\\ detected solar-mass cluster members plus the upper limit for the one undetected source is very likely to include all the cluster members with masses in the $0.8-1.2~M_{\\sun}$ range, in this field. The median luminosity value for this sample is $1.3 \\times 10^{28}$\\es; the quantiles bars also refer to this sample. Including the two candidate members with $M \\sim 0.8 M_{\\sun}$ (Ids 136 and 246 -- see Table\\,\\ref{tab:lx}) does not change the median value; including the one \\xmm\\ detected (solar-mass) member outside the \\chandra\\ FOV does not change this value either, although it increases the size of the 75\\% quantile bar. The error on the distance of the cluster ($430\\pm 20$ pc) is relatively small and does not impact significantly on the position of the median luminosity of NGC~752 in the plot; for the distance range of $410-450$ pc, the value of the median X-ray luminosity stays within the two quantile vertical bars. The X-ray luminosities of the seven points in NGC~752 have a spread of more than one order of magnitude, confirming the large spread in luminosity of stars with the same age and similar mass already observed in the Pleiades, the Hyades (e.g. \\citealp{ssk95}; \\citealp{msk+96}) and in the sample by \\cite{nb98} studied by \\cite{micela02}. The observed spread in $L_{\\rm X}$ at a fixed age is associated to the spread in rotational periods (\\citealp{pmm+03}), which appears to depend on the coupling between the circumstellar disc and the star in the pre-main sequence phase; in some objects this coupling prevents a young star from spinning up during its PMS contraction, yielding a large spread in the initial angular momentum distribution (e.g. \\citealp{bouvier94}; \\citealp{ch96}). The analysis of $\\sim$ 500 pre-main sequence and recently arrived main-sequence stars by \\cite{hm05} supports this view. The median X-ray luminosity of 1.3 $ \\times 10^{28}$\\es for solar-type stars in NGC~752 is in good agreement with the decaying trend of X-ray luminosity from the Hyades to the field stars. The value is about 6 times lower than the median value for the Hyades and 6.5 times higher than the median value from field stars, consistent with a steepening of the X-ray luminosity scaling law after the age of the Hyades. In a study of nine solar-like G-stars with ages ranging from 70 Myr to 9 Gyr, \\cite{ggs97} found $L_{\\rm X} \\propto t^{-\\beta}$ with $\\beta \\sim 1.5$, for stars with ages beyond a few 100 Myr, in agreement with the earlier results by \\cite{msv+87}. The study of a sample of 11 late-type stars in the Chandra Deep Field-North by \\cite{fhm+04} is also consistent with this scaling law, for $ 1 < t < 11$ Gyr, although an excellent fit to their data is found for $\\beta=2.0$. \\cite{pp04} find evidence for a very steep decay of the chromospheric activity between 0.5 and 2 Gyr, their sample at $\\sim 2$ Gyr consisting of seven stars. Comparison of the median X-ray luminosity of NGC~752 with that of the Hyades is fully consistent with a scaling law with $\\beta \\sim 1.5$. This scaling law is also consistent with the decay in X-ray luminosity from NGC~752 to the field stars, given the age uncertainties of the field stars. Since $L_{\\rm X} \\approx v_{\\rm rot}^{2}$ (e.g. \\citealp{pgr+81}; \\citealp{pmm+03}), this result implies $\\alpha \\sim 0.75$ for the scaling law of rotational velocities, for $t \\ga 1$ Gyr. This is significantly steeper than found by \\cite{skuma72} and would require a nearly-dipolar magnetic configuration to be explained in terms of magnetic breaking (\\citealp{kawaler88}). The effect of differential rotation in the stellar interior and the onset of magnetic saturation, however, may also play a role (\\citealp{kpb+97}) and it is possible that the coupling efficiency between outer and inner layers of stars weakens with age or that it is mass dependent (\\citealp{barnes03}). Comparison of rotational velocities in the Pleiades with those in older clusters such as M34 and the Hyades shows that, within the age interval of the Pleiades and Hyades, a star's rotational rate typically decreases less steeply than predicted by a pure magnetic braking law, that is $\\alpha < 0.3$ (\\citealp{qam+98}). The decay of the median $L_{\\rm X}$ of stars with mass $M=0.8-1.2~M_{\\sun}$ from the Pleiades to the Hyades reported by \\cite{micela02} (the points shown in Fig.\\,\\ref{fig:lxvage}) and that reported by \\cite{pf05} for stars with $M=0.9-1.2~M_{\\sun}$ for the same clusters, confirm this result ($\\alpha < 0.25$). As discussed by \\cite{qam+98}, a value of $\\alpha < 0.3$ could be an indication that angular momentum tapped in the radiative core of slow rotators on the zero age main sequence resurfaces into the convective envelope between the Pleiades and Hyades ages. We know from helioseismology that there is no gradient between the angular velocities of the core and the envelope in the Sun, thus our data suggest a change in rotation regimes of the stellar interior at $t\\sim $ 1 Gyr. The shape of the temporal evolution of the X-ray luminosity of solar-mass stars also has an important consequence in the evolution of close-in exoplanets $-$ within 0.5 AU. As shown by \\cite{pml08}, the later the timing of the transition between the two scaling laws of $L_{\\rm X}$, the smaller the fraction of gaseous planets which at 4.5 Gyr retain most of their initial mass. Our result indicates this transition to be at around 1 Gyr. \\begin{figure}[] \\begin{center} \\leavevmode \\epsfig{file=10042f10.ps, width=8.0cm} \\caption{X-ray luminosity of members of NGC~752 with mass between 0.8 and 1.2 $M_{\\sun}$ (filled points for \\chandra\\ detected stars and empty point for the one detected by \\xmm\\ outside \\chandra\\ FOV), compared with median luminosities of stars within the same mass range in the Pleiades, Hyades and a sample of field stars. The triangle indicates the upper limit for star Id 790 from DLM94. Error bars indicate the median value with the size of the vertical bar indicating the 25\\% and 75\\% quantiles of the distribution (see text). The line at age 4.5 Gyr connects the minimum and maximum of the Sun.} \\label{fig:lxvage} \\end{center} \\end{figure}" }, "0808/0808.4047_arXiv.txt": { "abstract": "This Letter is to investigate the physics of a newly discovered phenomenon --- contracting flare loops in the early phase of solar flares. In classical flare models, which were constructed based on the phenomenon of expansion of flare loops, an energy releasing site is put above flare loops. These models can predict that there is a vertical temperature gradient in the top of flare loops due to heat conduction and cooling effects. Therefore, the centroid of an X-ray looptop source at higher energy bands will be higher in altitude, for which we can define as normal temperature distribution. With observations made by {\\it RHESSI}, we analyzed 10 M- or X-class flares (9 limb flares). For all these flares, the movement of looptop sources shows an obvious U-shaped trajectory, which we take as the signature of contraction-to-expansion of flare loops. We find that, for all these flares, normal temperature distribution does exist, but only along the path of expansion. The temperature distribution along the path of contraction is abnormal, showing no spatial order at all. The result suggests that magnetic reconnection processes in the contraction and expansion phases of these solar flares are different. ", "introduction": "It is widely accepted that solar flare energy comes from sudden release of free magnetic energy via magnetic reconnection. In commonly adopted flare models, the ever-ascending Y-type reconnection point in the solar corona results in expanding flare loops and separation motion of flare foot points (FPs) (Kopp \\& Pneuman 1976). The expansion of flare ribbons and flare loops is an important signature of progressive magnetic reconnection in the corona. However, the contraction of flare loops in the early phase of flares may suggest a different reconnection scenario. The signature for the contraction of flare loops has been reported by several authors (Sui et al. 2003, 2004; Li \\& Gan 2005; Liu et al. 2004; Veronig et al. 2006; Ji et al. 2004b, 2006, 2007, 2008). Contraction motion includes two aspects: the converging motion of FPs and the correlated downward motion of a LT source, such as the M1.1 flare of 2004 November 1 (Ji et al. 2006). The contraction picture is different from the shrinkage of flare loops with rooted FPs being fixed or still in expansion. From our experience, the similar event like the M1.1 flare is rare, since HXR FPs are usually missing during the initial phase of a flare, such as the X3.9 flare of 2003 November 3 (Veronig et al. 2006). To investigate the converging motion of FPs, the most preferable wavelength is H$_\\alpha$ blue wing, at which H$_\\alpha$ emission is believed to be caused by nonthermal electrons (Canfield et al. 1984). The classical flare model puts the energy releasing site above flare loops, from which we can expect that a higher temperature source will be located above a lower temperature source due to heat conduction and cooling effects. In this Letter, we will name this distribution as normal temperature distribution (NTD). For the X10 flare of 2003 October 29, Ji et al. (2008) reported that a HXR sigmoid structure contracts during the impulsive phase of the flare. The contraction is the result of reconnection between two highly-sheared flux ropes. For magnetic reconnection between flux ropes, the existence of NTD is not required, since the energy releases inside flux ropes. Thus, temperature distribution along the altitude is an important factor for testing the reconnection scenarios. Based on above thinking, we re-analyzed the M2.1 flare of 2002 September 9. The flare is a well-observed sample showing the early converging and subsequent separation motion of the flare kernels. We found that the motion of the HXR LT source is well correlated with that of the flare kernels. Notably, the NTD occurs only during the expansion period of flare loops. During the contraction period, higher and lower temperature structures are mixed together. To find supporting evidences for this abnormal temperature distribution during the contraction period, we surveyed nearly all M-class and X-class limb flares from 2002 to 2005, which were well-observed by {\\it RHESSI} (Lin et al. 2002), we found at least 9 events showing the phenomenon. In \\S 2 we present the result of the flare of 2002 September 9, in \\S 3 we give the other 9 supporting events. Discussions and conclusions are briefly given in \\S 4. ", "conclusions": "In this Letter, we report the finding of the early abnormal temperature distribution of flares' X-ray LT sources during contraction phase of ten solar flares. In a 2D framework for flare models, an energy releasing site is assumed to be above flare loops. Following the scenario, we can expect that a LT source at higher energies will be higher in altitude due to heat conduction and cooling effects. The results from the ten flares show that this expectation can be met, but only during the expansion phase of solar flares. During the contraction period, however, high energy and low energy LT sources are mixed, showing a kind of abnormal temperature distribution. The physics for the contraction of flare loops is still not well-understood. From above ten events, we have seen that the turning points in the trajectories of the LT sources occur near flare peak times. Therefore, the contraction is related to magnetic reconnection in the impulsive phase of solar flares. Ji et al. (2008) reported that, during the contraction period of the X10 flare of 2003 October 29, HXR emissions at all energies share the similar sigmoidal configuration. The contraction corresponds to the shrinkage of the HXR sigmoid, which is the result of magnetic reconnection between highly-sheared flux ropes. We may propose that the abnormal temperature distribution is associated with magnetic reconnection between highly-sheared flux ropes. On the other hand, the downward and upward phases of the LT motion may reflect two regimes of reconnection: bursting and no-bursting (Karlick\\'{y} 2008: private communication). The abnormal temperature structure can be explained by a presence of small plasmoids in the current sheet just above contracting arcade. The plasmoids are formed in the bursting regime of the reconnection and move also downwards where they interact with the arcade. This interaction represents additional reconnection that changes the normal temperature structure (B\\'{a}rta et al. 2008; Koloma\\'{n}ski \\& Karlick\\'{y} 2007). In the phase of the upward LT motion the reconnection take place without these plasmoids (no-bursting regime of the reconnection). Therefore the temperature is in agreement with the prediction made by standard flare models." }, "0808/0808.3983_arXiv.txt": { "abstract": "We present multi-color optical observations of long-duration $\\gamma$-ray bursts (GRBs) made over a three year period with the robotic Palomar 60\\,inch telescope (P60). Our sample consists of all 29 events discovered by \\Swift\\ for which P60 began observations less than one hour after the burst trigger. We were able to recover $80\\%$ of the optical afterglows from this prompt sample, and we attribute this high efficiency to our red coverage. Like \\citet{mmk+08}, we find that a significant fraction ($\\approx 50\\%$) of \\Swift\\ events show a suppression of the optical flux with regards to the X-ray emission (so-called ``dark'' bursts). Our multi-color photometry demonstrates this is likely due in large part to extinction in the host galaxy. We argue that previous studies, by selecting only the brightest and best-sampled optical afterglows, have significantly underestimated the amount of dust present in typical GRB environments. ", "introduction": "\\label{sec:intro} The launch of the \\Swift\\ $\\gamma$-Ray Burst (GRB) Explorer \\citep{gcg+04} in 2004 November has ushered in a new era in the study of GRB afterglows. \\Swift\\ offers a unique combination of event rate ($\\sim 100$\\,yr$^{-1}$; almost an order of magnitude increase over previous missions) and precise localization ($\\sim 3\\arcmin$ radius error circles are distributed seconds after the burst, and refined to $\\sim 3\\arcsec$ minutes later). The on-board X-ray Telescope (XRT; \\citealt{bhn+05}) and UV-Optical Telescope (UVOT; \\citealt{rkm+05}), together with the rapid relay of these precise localizations to ground-based observers, has enabled an unprecedented glimpse into the time period immediately following the prompt emission over a broad frequency range. Observations of X-ray afterglows with the XRT have generated particular interest in recent years. In the pre-\\Swift\\ era, X-ray observations were limited to hours or days after the prompt emission, and were often poorly sampled compared with the optical and radio bandpasses. Routine XRT observations of \\Swift\\ GRBs beginning at early times have revealed a central engine capable of injecting energy into the forward shock at times well beyond the duration of the prompt emission (e.g., \\citealt{brf+05,zfd+06}). This discovery has had a profound effect on our understanding of progenitor models. While the X-ray afterglow is currently a well-explored phase space, comparatively few analogous studies have been performed in the optical bandpass. \\citet{bkf+05} first suggested that \\Swift\\ optical afterglows were 1.8\\,mag fainter in the $R$-band than pre-\\Swift\\ events (at a common epoch of 12\\,hours after the burst). Likewise, \\citet{rsf+06} found that only 6 of the first 19 \\Swift\\ bursts with prompt ($\\Delta t \\lesssim 100$\\,s) UVOT coverage yielded optical afterglow detections. Since then, explaining the faintness of \\Swift\\ optical afterglows has remained one of the outstanding questions in the field. One clear contributor is distance: the median redshift of \\Swift\\ events ($\\langle z_{Swift} \\rangle \\approx 2.0$)\\footnote{Calculated from J.~Greiner's compilation at http://www.mpe.mpg.de/\\~{}jcg/grbgen.html.} is significantly larger than the pre-\\Swift\\ sample ($\\langle z_{\\mathrm{pre-}Swift} \\rangle = 1.1$; \\citealt{bkf+05,jlf+06}). In a comprehensive literature-based study of the brightest, best-studied \\Swift\\ afterglows, \\citet{kkz+07} find properties broadly similar to pre-\\Swift\\ events, after applying a cosmological k-correction. On the other hand, \\citet{mmk+08} have recently presented a sample of 63 GRBs observed in the optical ($r^{\\prime}$-band) with the robotic 2\\,m Liverpool Telescope and Faulkes Telescopes (North and South). The selection criteria for including a burst in their sample is never explicitly stated, and several non-\\Swift\\ bursts are included, making a direct comparison with the results of \\citet{kkz+07} difficult. However, \\citet{mmk+08} do not exclude the significant fraction of events without optical detections from their analysis, providing a more unbiased look at optical afterglow properties. By measuring the ratio of optical to X-ray flux at a common time, these authors find that roughly half of the GRBs in their sample exhibit a relative suppression of the optical flux inconsistent with our standard picture of afterglow emission (e.g., \\citealt{spn98}), so-called ``dark'' bursts \\citep{jhf+04}. This finding suggests that distance alone cannot explain the faintness of \\Swift\\ optical afterglows. Several other possibilities have been suggested to explain optically dark GRB afterglows. Undoubtedly some GRBs, like GRB\\,050904 \\citep{hnr+06,kka+06}, originate from such large redshifts ($z \\gtrsim 6$) that Ly-$\\alpha$ absorption in the inter-galactic medium (IGM) completely suppresses the optical flux \\citep{lr00}. Alternatively, late-time energy injection from the central engine, manifested as bright X-ray flares and/or extended periods of shallow decay, may be artificially increasing the X-ray flux, leading to spurious claims of optically dark GRBs \\citep{mmk+08}. One final possibility is extinction native to the GRB host galaxy. As a population, long-duration GRB host galaxies exhibit extremely large neutral H column densities (e.g., \\citealt{hmg+03,bpc+06}), typically falling at $\\log N_{H} > 20.3$\\,cm$^{-2}$ (so-called Damped Ly-$\\alpha$, or DLA systems; \\citealt{wgp05}). And within their hosts, GRBs trace the blue light from hot young stars in the disk even more closely than core-collapse supernovae \\citep{bkd02,fls+06}. Both findings are consistent with the observed association between long-duration GRBs and massive star death (e.g., \\citealt{wb06}). In spite of these expectations, relatively few GRB afterglows to date exhibit signs of large host galaxy extinction (e.g., \\citealt{cbm+07,rvw+07,tlr+08}). \\citet{kkz+07} find only a modest amount of dust ($\\langle A_{V} \\rangle = 0.20$\\,mag) for the 15 events in their ``golden'' sample, an identical value found from an analogous study of pre-\\Swift\\ afterglows \\citep{kkz06}. The primary drawback of such studies, however, is the large and uncertain role of selection effects: by including only the brightest, best-sampled optical afterglows, \\citet{kkz+07} may be preferentially selecting those events in low-extinction environments. Understanding these selection effects is one of the primary goals of this work. The Palomar 60\\,inch telescope (P60) is a robotic, queue-scheduled facility dedicated to rapid-response observations of GRBs and other transient events \\citep{cfm+06}. With a response time of $\\Delta t \\lesssim 3$\\,min and a limiting magnitude of $R \\gtrsim 20.5$ (60\\,s exposure), the P60 aperture is well suited to detect most \\Swift\\ optical afterglows \\citep{as07}. In addition, with a broadband filter wheel providing coverage from the near-UV to the near-IR, P60 can also provide multi-color data on the afterglow evolution. In this work, we present the P60-\\Swift\\ Early Optical Afterglow sample: 29 unambiguously long-duration GRBs detected by the \\Swift\\ Burst Alert Telescope (BAT; \\citealt{bbc+05}) with P60 observations beginning at most one hour after the burst trigger time. This sample offers two distinct advantages over previous efforts to understand the optical afterglow emission from GRBs. First and foremost, our study enforces a strict selection criterion independent of the optical afterglow properties, and therefore will allow us to study the properties of the \\Swift\\ population in a relatively unbiased manner. Secondly, nearly all events contain multi-color ($\\gp\\,\\Rc\\,\\ip\\,\\zp$) observations that allow us to evaluate the importance of host galaxy extinction for a fraction of our sample. Altogether, we aim to discriminate between the competing hypotheses proffered to explain dark GRB afterglows in the \\Swift\\ era. Throughout this work, we adopt a standard $\\Lambda$CDM cosmology with $h_{0}$ = 0.71\\,km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{\\mathrm{m}} = 0.27$, and $\\Omega_{\\Lambda} = 1 - \\Omega_{\\mathrm{m}} = 0.73$ \\citep{sbd+07}. We define the flux density power-law temporal and spectral decay indices $\\alpha$ and $\\beta$ as $f_{\\nu} \\propto t^{-\\alpha} \\nu^{-\\beta}$ (e.g., \\citealt{spn98}). All errors quoted are 1 $\\sigma$ (i.e., $68\\%$) confidence intervals unless otherwise noted. ", "conclusions": "\\label{sec:discussion} \\subsection{Anomalous P60 Detection Efficiency} \\label{sec:deteff:disc} We have demonstrated in \\S~\\ref{sec:deteff:obs} that P60 was able to detect optical afterglow emission from a large fraction ($\\sim 80\\%$) of events for which observations began within an hour of the burst trigger. While the 1.5\\,m aperture is relatively large for a robotic facility, it would be nonetheless informative to understand systematic effects that affect our afterglow recovery rate. The ultimate goal, of course, is to better inform future GRB follow-up campaigns. The first lesson from this campaign is the importance of observing in redder filters. We have shown in \\S~\\ref{sec:dark:obs} that typical \\Swift\\ events suffer from a non-negligible amount of host galaxy extinction (Tab.~\\ref{tab:early}). Coupled with the additional effect of Ly-$\\alpha$ absorption in the IGM from a median redshift of $\\langle z \\rangle \\approx 2$, it is clear that a large fraction of the low UVOT detection efficiency is caused by its blue observing bandpass. The P60 automated follow-up sequence, consisting of alternating exposures in the \\Rc, \\ip, and \\zp\\ filters, while initially designed for identification of candidate high-$z$ events, is actually well-suited to maximize afterglow detection rates. The large fraction of P60-detected bursts with spectroscopic redshifts, on the other hand, is almost certainly an artifact of the unequal longitudinal distribution of large optical telescopes. With the exception of the South African Large Telescope (SALT), all optical telescopes with apertures larger than 8\\,m fall within six time zones (UT-4 to UT-10). It is not entirely surprising then, that so many promptly discovered P60 optical afterglows have spectroscopic redshifts from immediate follow-up with the largest optical facilities. While building the largest optical facilities is often prohibitively expensive for all but the largest collaborations, 1\\,m class facilities are much more feasible, both in terms of cost and construction time scale. We wish here to echo the thoughts of many previous GRB observers (e.g., \\citealt{as07}) that future automated facilities be built at longitudes (and latitudes) not covered by current facilities. NIR coverage is particularly crucial to detect the most extinguished events and provide tighter constraints on the afterglow SED and hence host galaxy extinction. A longitudinally spaced ring of 1\\,m class facilities, as for example envisioned by the Las Cumbres Observatory Global Telescope\\footnote{See http://lcogt.net.} is well positioned in the future to recover the vast majority of GRB optical afterglows, assuming the follow-up is done in the reddest filters possible. Such coverage will be particularly important as we transition into the \\textit{Fermi} era, with its significantly decreased rate of precise GRB localizations. \\subsection{Re-visiting Dark Bursts} \\label{sec:darkbursts:disc} We now turn our attention to the issue of dark bursts in the \\Swift\\ era. In \\S~\\ref{sec:dark:obs}, we demonstrated that a large fraction ($\\approx 50\\%$) of \\Swift\\ afterglows showed suppressed emission in the optical bandpass (relative to the X-ray), that was due in large part to extinction in the host galaxy. Given the natural expectation that GRBs, since they are associated with massive stars, should form in relatively dusty environments, we wish to understand why our study of \\Swift\\ events yields such a dramatically different dark burst fraction than previous work on pre-\\Swift\\ GRBs \\citep{jhf+04}. We believe selection effects are one large cause of this discrepancy. It is clear that previous studies of GRB host galaxy extinction, by selecting the brightest and best-sampled events, provide a strongly biased view. Many, if not most, GRB hosts, appear to suffer from a significant amount of dust extinction ($A_{V} \\gtrsim 0.5$). Even the study of \\citet{jhf+04}, though it included \\textit{all} pre-\\Swift\\ GRBs with an X-ray afterglow, could be biased towards unextinguished events as well. Before \\Swift, target-of-opportunity X-ray observations often required the accurate localization provided by an optical (or radio) afterglow. Thus those events with the brightest optical afterglows (assumed to have on average smaller extinction) were more likely to be observed in the X-ray, biasing the optical-to-X-ray spectral index to larger values of $\\beta_{OX}$. Another, more subtle, effect, may also cause \\Swift\\ afterglows to appear darker than pre-\\Swift\\ afterglows, independent of host galaxy extinction. Because \\Swift\\ is a more sensitive instrument, it detects GRBs at a higher average redshift than any previous mission. Consider a host frame extinction of $A_{V} = 0.1$\\,mag. At $z = 1$, typical for pre-\\Swift\\ events, the observed \\Rc\\ filter corresponds to roughly to rest-frame $U$-band, and so an extinction of 0.17\\,mag (assuming a Milky Way-like extinction curve). On the other hand, at $z = 3$, the observed \\Rc-band corresponds to a rest frame wavelength of $\\lambda = 1647$\\,\\AA. So at high redshift, the same amount of dust will produce nearly twice as much extinction in the observed bandpass. Solely because of redshifts effects, similar environments will produce different observed spectral slopes. This effect is exacerbated by the nature of dust grains in most GRB host galaxies, as the SMC extinction curve does not show the pronounced 2175\\,\\AA\\ bump seen from the Milky Way \\citep{p92}. If GRBs do trace the cosmic star formation rate, our results suggest a significant fraction of star formation occurs in highly obscured environments. \\citet{kkz06} found a weak correlation between host reddening and sub-mm flux, and we believe a sensitive mid-infrared or sub-mm survey of GRB host galaxies would be an important confirmation of our results. However, instead of focusing on the brightest, best studied afterglows, as has often been done in the past (e.g., \\citealt{tbb+04,mhc+08}), we instead suggest a survey of the host galaxies of the optically darkest GRB afterglows to see if these events really do exhibit signs of obscured star formation." }, "0808/0808.2917_arXiv.txt": { "abstract": "Gamma-ray bursts (GRBs) are cosmologically distributed, very energetic and very transient sources detected in the $\\gamma$-ray domain. The identification of their $x$-ray and optical afterglows allowed so far the redshift measurement of \\ngrbs\\ events, from $z=0.01$ to $z=6.29$. For about half of them, we have some knowledge of the properties of the parent galaxy. At high redshift ($z>2$), absorption lines in the afterglow spectra give information on the cold interstellar medium in the host. At low redshift ($z<1.0$) multi-band optical-NIR photometry and integrated spectroscopy reveal the GRB host general properties. A redshift evolution of metallicity is not noticeable in the whole sample. The typical value is a few times lower than solar. The mean host stellar mass is similar to that of the Large Magellanic Cloud, but the mean star formation rate is five times higher. GRBs are discovered with $\\gamma$-ray, not optical or NIR, instruments. Their hosts do not suffer from the same selection biases of typical galaxy surveys. Therefore, they might represent a fair sample of the most common galaxies that existed in the past history of the universe, and can be used to better understand galaxy formation and evolution. ", "introduction": "\\begin{figure} \\centerline{\\includegraphics[scale = .4]{fig1.eps}} \\begin{center} \\caption{Histogram of GRBs with measured redshift (empty histogram \\ngrbs\\ objects). The high-$z$ and low-$z$ filled histograms are the subsample of GRBs studied using optical afterglow spectra (GRB-DLAs) and multiband photometry of the host galaxies (GRB hosts), respectively.} \\label{z_hist} \\end{center} \\end{figure} More than 40 years have passed since the discovery of the first gamma-ray burst (GRB), the most energetic explosions in the universe, by the US military satellite Vela. It took 30 years to demonstrate their cosmological origin (\\cite[Metzger et al.\\ 1997]{metzger97}). Their $\\gamma$-ray energies (typically $10^{51}$ ergs, emitted in less than a couple of minutes) emerge from a collimated jet in a core-collapse supernovae, or the merger of two compact objects (neutron stars or black holes). Due to the highly transient nature of GRBs, their redshift, measured from the optical afterglow or the host galaxy, is known today for \\ngrbs\\ objects only (Figure~\\ref{z_hist}). Although GRBs are very rare (a rate of 1 event every $10^5$ years in a galaxy is estimated, after correcting for the jet opening angle), they are so energetic that a few events are expected to be detectable from Earth every day. As shown by the discovery of high redshift events (the highest ever being GRB~050904 at $z=6.3$; Kawai et al.\\ 2006), GRBs offer the opportunity to explore the most remote universe, under extreme conditions, hard to observe using traditional tools. The first scientific paper on GRBs dates from 6 years after the Vela discovery (Klebesadel, Strong \\& Olson 1973). In 1975, already 100 different theories where proposed, and today more than 5000 refereed papers on GRBs have been published. Before the 1997 discovery, theories on the energetic emission ranked from the impact of comets onto neutron stars (Harwit \\& Salpeter 1973) to collisions of chunks of antimatter with normal stars (Sofia \\& van Horn 1974). However, the most likely hypothesis was already postulated years earlier, when nothing was really known about GRBs: Colgate (1968) predicted $\\gamma$-ray emission from supernovae in distant galaxies. The curse and blessing of GRBs is their fast fading. The light curve is extremely steep and therefore very hard to catch. The emitted energy is so immense that it cannot last long. However, it allows us to see a hidden universe. One extreme case is the recent event GRB~080319B at $z=0.937$ (Bloom et al.\\ 2008), visible for a short time by naked eye (pick optical magnitude $m=5.6$). This GRB was already hard to observe spectroscopically with the largest telescopes 7 hours after its discovery. \\begin{table} \\begin{center} \\caption{Overview of host diagnostics with GRB observations.} \\label{t1} {\\scriptsize \\begin{tabular}{|l|c|c|c|c|}\\hline & {\\bf GRB-DLA} & {\\bf GRB host} & {\\bf GRB host} \\\\ {\\bf Diagnosis} & {\\bf (AG spectroscopy)} & {\\bf (Int. spectroscopy)} & {\\bf (Photometry)} \\\\ \\hline SFR & $\\times$ & $\\surd\\surd$ & $\\surd$ \\\\ \\hline Metallicity & $\\surd\\surd\\surd$ & $\\surd\\surd$ & $\\times$ \\\\ \\hline Dust extinction & $\\surd$ & $\\surd\\surd$& $\\times$ \\\\ \\hline Dust depletion & $\\surd\\surd\\surd$ & $\\times$ & $\\times$ \\\\ \\hline Stellar mass & $\\times$ & $\\times$ & $\\surd\\surd\\surd$ \\\\ \\hline Age & $\\times$ & $\\surd$ & $\\surd$ \\\\ \\hline \\end{tabular} } \\end{center} \\end{table} ", "conclusions": "The main, still unsolved, question about GRB hosts is whether they represent a fair sample of the whole star-forming galaxy population, or they are a distinct population of galaxies. GRB hosts are generally small, star-forming galaxies detected at any redshift, up to $z=6.3$. The mean stellar mass (measured mainly for the low-$z$ subsample) is a few times above $10^9$ M$_\\odot$. That is the stellar mass of the Large Magellanic Cloud. The mean SFR is 5 times higher than the LMC (Savaglio et al.\\ 2008). Metallicities of the hosts, measured from absorption lines in the afterglow spectra at $z>2$ or from emission lines in the host spectra at $z<1.0$, do not indicate a clear redshift evolution (Figure~\\ref{fig4}), with values mainly between solar and 1/10 solar. This suggests that the chemical enrichment is not a parameter characterizing the GRB host population. It is clear that small star-forming galaxies are the most common galaxies that existed in the entire history of the universe. For instance, the star formation history of the universe, studied for different galaxy stellar-mass bins, indicates that small galaxies dominated for redshift $z<2$, where massive galaxies experienced a fast decline (Juneau et al.\\ 2005). GRB hosts are low stellar-mass star-forming galaxies and can provide a very efficient way to understand galaxy formation and evolution in the most active phase of the universe. Future missions will give the possibility to study them in much larger quantities. \\begin{acknowledgement} We thank the conference organizers for the outstanding organization. The author thanks Karl Glazebrook and Damien Le Borgne for the very fruitful and long-lasting collaboration. \\end{acknowledgement}" }, "0808/0808.1639_arXiv.txt": { "abstract": "Orbital periods in AM Her stars (\\emph{polars}) are synchronized with spin periods of white dwarf by its high magnetic field. Since the last study of $P_{orb}$ distribution of these systems, the number of known objects of such type has more than doubled. This challenged us to compile a new updated catalogue of cataclysmic variables with highly magnetic white dwarfs (polars) and to study their $P_{orb}$ distribution. In this paper we also discus if \"spike\" is reliable feature in the distribution. (\"Spike\" is a concentration of polars in the distribution of their orbital periods near $P_{orb}$ = 114 min and was previously discussed by Ritter \\& Kolb (1992) and Shahbaz \\& Wood (1996).) ", "introduction": "AM Her stars or Polars - is a subtype of cataclysmic variables, where binary stellar system contains highly magnetic white dwarf. The strength of magnetic field allows to control accretion in such system (mass is transferring directly on the white dwarf pole, without formation of accretion disc). As a distinct from intermediate polars, AM Her stars orbital period is synchronized with white dwarf spin period. Last study of $P_{orb}$ distribution were made by Shahbaz \\& Wood (1996) in 1996 (just 43 systems were known by that time). They have noted that discovery of one more AM Her-type star with the orbital period outside of the spike will decrease its significance below 99\\% level. Since that time the number of known systems has more then doubled. This incentivised us to compile a new updated catalogue of cataclysmic variables with highly magnetic white dwarfs and to study distribution of their orbital periods and to re-calculate significance of spike in the similar way, as it was done in previous work. ", "conclusions": "\\begin{itemize} \\item Most of known polars have orbital periods below \"period gap\" (50 of 91 known polars). \\item We obtained a grater statistic (91 systems in total) and were impelled to make the spike ranges some broader: $\\Delta P_{spike}$ = 4 min ($\\Delta P_{spike}$ = 2 min in Ritter \\& Kolb (1992) - 17 polars in total were known, $\\Delta P_{spike}$ = 3 min in Shahbaz \\& Wood (1996) - 43 polars were known). \\item Though spike remains to be significant feature, its significance has decreased and now its value is about $99.6 \\%$, instead of the 99.9\\%, as it was in the work of T.Shahbaz \\& Janet H.Wood (1996). \\item The observed number of systems in the gap is inconsistent with the period distribution being uniform at the 96.3\\% level. So we have a lowering of this value, just like as for the spike significance. \\end{itemize}" }, "0808/0808.1580_arXiv.txt": { "abstract": "We generalize the Bekenstein-Sandvik-Barrow-Magueijo (BSBM) model for the variation of the fine structure 'constant', $\\alpha ,$ to include an exponential or inverse power-law self-potential for the scalar field $% \\varphi $ which drives the time variation of $\\alpha $, and consider the dynamics of $\\varphi $ in such models. We find solutions for the evolution of $\\varphi $ or $\\alpha $ in matter-, radiation- and dark-energy-dominated cosmic eras. In general, the evolution of $\\varphi $ is well determined solely by either the self-potential or the coupling to matter, depending on the model parameters. The results are general and applicable to other models where the evolution of a scalar field is governed by a matter coupling and a self-potential. We find that the existing astronomical data stringently constrains the possible evolution of $\\alpha $ between redshifts $z\\simeq 1-3.5$ and the present, and this leads to very strong limit on the allowed deviation of the potential from that of a pure cosmological constant. ", "introduction": "\\label{sect:Introduction} For the first time there is a body of detailed astronomical evidence consistent with the time variation of a traditional constant of Nature. The observational programme of Webb \\emph{et al}. \\cite{webb1,webb2} has completed detailed analyses of 128 Keck-HIRES quasar absorption line systems at redshifts $0.54.5\\times 10^{-8}$ $(6\\sigma ),$ when the non-thermal neutron spectrum is taken into account. However, there remain significant environmental uncertainties regarding the reactor's early history and the relationship between changes in the resonance energy level and those in the values of any underlying constants. For reviews of the wider issue of varying constants in addition to $\\alpha $, see the reviews in refs.~\\cite{revs}, and for some implications of the unification of fundamental forces see refs.~\\cite{unify}. Recently, Rosenband \\emph{et al}. \\cite{Rosenband} measured the ratio of aluminium and mercury single-ion optical clock frequencies, $f_{\\mathrm{Al+}% }/f_{\\mathrm{Hg+}}$, at intervals over a period of about a year. From these measurements, the linear rate of change in this ratio was found to be $% (-5.3\\pm 7.9)\\times 10^{-17}\\,yr^{-1}$ (but see ref. \\cite{bs} for some refinements). These measurements provides the strongest limit yet on any temporal drift in the value of $\\alpha $: \\begin{equation*} \\ \\dot{\\alpha}/\\alpha =(-1.6\\pm 2.3)\\times 10^{-17}\\,yr^{-1}.\\ \\end{equation*}% This limit is strong enough to exclude theoretical explanations of the change in $\\alpha $ reported by Webb \\emph{et al}. \\cite{webb1, webb2} based on the slow variation of an effectively massless scalar field \\cite{bsbm}, even allowing for the damping by cosmological acceleration, unless there is a significant effect that slows the locally observed effects of changing $% \\alpha $ on cosmological scales (for a detailed analysis of global-local coupling of variations in constants (see Refs.~\\cite{shawb}). Theories in which $\\alpha $ varies will in general lead to violations of the weak equivalence principle (WEP). This is because the $\\alpha $ variation is carried by a scalar field, $\\varphi ,$ and this couples differently to different nuclei because they contain different numbers of electrically charged particles (protons). The theory discussed here has the interesting consequence of leading to a relative acceleration of order $10^{-13}$ \\cite% {bmswep} if the free coupling parameter is fixed to the value given in Eq.~(% \\ref{om}) using a best fit of the theories cosmological model to the quasar observations of refs.~\\cite{webb1, webb2}. Other predictions of WEP violations have also been made in refs. \\cite{poly, zal, dam}. The observational upper bound on this parameter from direct experiment is just an order of magnitude larger, at $10^{-12},$ and limits from the motion of the Moon are of similar order, \\cite{nord}, but space-based tests planned for the STEP mission are expected to achieve a sensitivity of order $% 10^{-18} $ and will provide a completely independent check on theories of time-varying $e$ and $\\alpha $ \\cite{wep, step}. In view of this tension between direct local measurements and astronomical measurements of the fine structure 'constant' it is important to explore the widest possible range of self-consistent theoretical models for the time-evolution of $\\alpha $ so as to understand the possible evolutions of $% \\Delta \\alpha /\\alpha $ over the range $0$ 1 (i.e. n $<$ 1) S\\'ersic indices. It is tempting to suggest that such N index for the outer component might be linked to a truncation of the profile \\citep*{E08}, due to a possible interaction with NGC\\,3258. However, such hypothesis should be studied in the light of the confirmed membership status of FS90\\,110. Furthermore, it would be interesting to test if the outer component of confirmed cE galaxies displaying two components in their brightness profiles, could be fitted by a S\\'ersic law with N $>$ 1 (n $<$ 1). The variation of the ellipticity and position angle vs. radius displayed by FS90\\,110, resemble those found in M32. These changes occur at a galactocentric radius at which the outer component seems to begin to dominate, in agreement with what \\citet{G02} found for M32. FS90\\,110 is the only FS90 cE candidate that follows the luminosity vs. mean effective surface brightness relation defined by bright ellipticals, which corresponds to the \\citet{K77} scaling relation, towards fainter magnitudes, smaller radius, or higher mean effective surface brightness. It also seems to follow the Antlia members' colour-magnitude relation, though it is located at the red border. The elongation and twisting of FS90\\,110's outermost isophotes in the direction of NGC\\,3258 are noticeable. That would be consistent with the extremely low surface brightness structure detected in the MOSAIC images, and confirmed with the FORS1 and ACS frames. Such kind of `bridge' that seems to link FS90\\,110 with its bright partner, would be in agreement with the fact that most confirmed cE galaxies are located in the vicinity of brighter companions. In that sense, it is also remarkable the warped inner structure displayed in the ACS colour map, and confirmed through unsharp masks of FORS1. Quite similar stellar structures are found in numerical simulations \\citep{B01,M06} as a consequence of galaxy interactions. \\subsection{Final remarks} A galaxy that is interacting with a more massive partner will feel tidal forces most strongly in its outskirts, while its central region will be less affected \\citep[see, for example,][for a model of the interaction between M32 and M31]{B01}. As a consequence, it would not be surprising to observe asymmetric and `egg-shaped' outer isophotes, or/and detect low surface brightness stellar `tails', arising from the disruption of the outer regions of the satellite \\citep{M06}. In particular, these faint structures could be observed either like one or two symmetric `tails', depending on projection effects \\citep[see fig.\\,4 in][]{M06}. The distorted outermost isophotes of FS90\\,110, and the extremely faint structure detected in the HST, VLT and CTIO images fit quite well in such a scenario. The similarities in the ellipticity and position angle variations against radius displayed by FS90\\,110 and FS90\\,208, which arise at a similar galactocentric radius of $\\sim 5.5$ arcsec (i.e. in their inner regions), are also remarkable. These kind of variations have already been detected in M32, although not so strongly. All these pieces of evidence lead us to speculate about the possibility that, an object similar to FS90\\,208 might be the progenitor of a cE galaxy. Thus, it is tempting to look for possible links between these two kind of objects. We may think that a system with similar characteristics as FS90\\,208 may lose its outermost regions due to the interaction with a bright companion, then becoming compact. As a consequence, it could also experience a redistribution of its stellar content, that might produce a warped inner structure, a higher central surface brightness, and an attenuation of its ellipticity and position angle variations with radius. Moreover, the bulge-to-disc ratio in such object could increase after losing its outer parts due to the interaction. With regard to this point, it should be noted that the inner component of FS90\\,110 seems to be 3 times brighter than the outer one, while that of FS90\\,208 is just 1.6 times brighter in relation with the outer component. As an aditional point, we recall that the Antlia cluster seems to be particularly rich in S0 galaxies and FS90\\,208 seems to be one of them. If the evolutionary path of cE galaxies include this kind of objects, S0 rich clusters would arise as favorable environments for the formation of compact galaxies. Dynamical simulations and the spectroscopically confirmed membership status of FS90\\,110 and FS90\\,192, will help to test if any of the above statements is indeed plausible." }, "0808/0808.0190_arXiv.txt": { "abstract": "We show that simple scalar field models can give rise to curvature singularities in the effective Friedmann dynamics of Loop Quantum Cosmology (LQC). We find singular solutions for spatially flat Friedmann-Robertson-Walker cosmologies with a canonical scalar field and a negative exponential potential, or with a phantom scalar field and a positive potential. While LQC avoids big bang or big rip type singularities, we find sudden singularities where the Hubble rate is bounded, but the Ricci curvature scalar diverges. We conclude that the effective equations of LQC are not in themselves sufficient to avoid the occurrence of curvature singularities. ", "introduction": " ", "conclusions": "" }, "0808/0808.1196_arXiv.txt": { "abstract": "Elucidating the processes that governed the assembly and evolution of galaxies over cosmic time is one of the main objectives of all of the proposed Extremely Large Telescopes (ELT). To make a leap forward in our understanding of these processes, an ELT will want to take advantage of Multi-Objects Adaptive Optics (MOAO) systems, which can substantially improve the natural seeing over a wide field of view. We have developed an end-to-end simulation to specify the science requirements of a MOAO-fed integral field spectrograph on either an 8m or 42m telescope. Our simulations re-scales observations of local galaxies or results from numerical simulations of disk or interacting galaxies. The code is flexible in that it allows us to explore a wide range of instrumental parameters such as encircled energy (EE), pixel size, spectral resolution, etc. For the current analysis, we limit ourselves to a local disk galaxy which exhibits simple rotation and a simulation of a merger. While the number of simulations is limited, we have attempted to generalize our results by introducing the simple concepts of ``PSF contrast'' which is the amount of light polluting adjacent spectra which we find drives the smallest EE at a given spatial scale. The choice of the spatial sampling is driven by the ``scale-coupling''. By scale-coupling we mean the relationship between the IFU pixel scale and the size of the features that need to be recovered by 3D spectroscopy in order to understand the nature of the galaxy and its substructure. Because the dynamical nature of galaxies are mostly reflected in their large-scale motions, a relatively coarse spatial resolution is enough to distinguish between a rotating disk and a major merger. Although we used a limited number of morpho-kinematic cases, our simulations suggest that, on a 42m telescope, the choice of an IFU pixel scale of 50-75 mas seems to be sufficient. Such a coarse sampling has the benefit of lowering the exposure time to reach a specific signal-to-noise as well as relaxing the performance of the MOAO system. On the other hand, recovering the full 2D-kinematics of z$\\sim$4 galaxies requires high signal-to-noise and at least an EE of 34\\% in 150 mas (2 pixels of 75 mas). Finally, we carried out a similar study for a hypothetical galaxy/merger at z=1.6 with a MOAO-fed spectrograph for an 8m, and find that at least an EE of 30\\% at 0.25 arcsec spatial sampling is required to understand the nature of disks and mergers. ", "introduction": "Developing a coherent model for the mass assembly of galaxies over cosmic time is a complex and difficult task. It is difficult because developing such a coherent model involves highly non-linear physics (e.g., the cooling and collapse of baryons, or feedback from stars and super-massive blackholes which regulate both the growth of the stellar mass and black holes themselves) and all over a very wide range of physical scales (from large scale structure to star clusters to black holes). Our knowledge of galaxy formation {\\it in situ} relies mostly on observations of integrated quantities,which is insufficient to understand the details of the process of formation, e.g., the interplay between the baryonic and dark matter, or the complex physics of baryons such as merging, star formation, feedback, various instabilities and heating/cooling mechanisms, etc. At the heart of our developing understanding of galaxy evolution is the relative importance of secular evolution through (quasi-)adiabatic accretion of mass, instabilities, and resonances (e.g., \\citealt{Semelin05}), versus more violent evolution through merging (e.g., \\citealt{Hammer05}), as a function of look-back time. Because of the complexity of the processes involved, understanding the mass assembly of galaxies over cosmic time requires constraining a wide range of phenomenology for large samples of objects. In this respect integral field spectroscopy is particularly well suited in that it allows us to map the spatially resolved physical properties of galaxies (see, e.g., \\citealt{Puech06b}). Unfortunately, splitting the light into many channels in a integral field spectrograph requires long integration times, even on large telescopes, to detect low surface brightness emission. Thus, conducting large statistical studies necessary for understanding the processes driving galaxy evolution will demand a high multiplex capability to efficiently investigate sufficient numbers of objects. Using current facilities on 8m class telescopes, it is already possible to map the physical properties of distant galaxies. For example, FLAMES/GIRAFFE on the VLT offers the unique ability to observe 15 distant galaxies simultaneously. The first (complete) 3D sample of galaxies at moderate redshifts has revealed a large fraction ($\\sim$40\\%) of all z$\\sim$0.6 galaxies with $M_B <$-19.5 have perturbed or complex kinematics (\\citealt{Flores06, Yang07}). This population of dynamically disturbed galaxies is most probably a result of a high prevalence of mergers and merger remnants (see also \\citealt{Puech06a, Puech08}). At higher redshift, integral field spectroscopy of several objects has been obtained using single object integral field spectrographs available on 8-10 meter class telescopes (e.g., \\citealt{Forster06}). However, many of these data sets were obtained with limited spatial resolution sampling galaxies on scales of a few kpc, and the true dynamical nature of many sources sampled at these spatial and spectral resolutions remains uncertain (e.g., \\citealt{Law07}). Improving and extending this approach to the general understanding of the nature of faint galaxies and to galaxies in the early Universe will require a new generation of multi-object integral field spectrographs working in the near-IR, with increased sensitivity and better spatial resolution on even larger telescopes. Because of the complex interplay between spatial and spectral features and our limited knowledge of high redshift galaxies, it is helpful, perhaps necessary, to rely on numerical simulations for constraining the design parameters of this new kind of instrument. For this purpose, we have developed software that simulate end-to-end the emission line characteristics of local galaxies and numerical simulations of galaxies to show how they would appear in the distant Universe. By changing various parameters like the resolution, pixel scale, and point spread function it is possible to constrain the instrumental characteristics and performance against a set of galaxy characteristics (e.g., velocity field). This paper is the first in a series to investigate the scientific design requirements of possible future instruments, telescope aperture and design, and site characteristics appropriate for constrain how galaxies grew and evolved with cosmic time. In this paper in particular, we aim to present the details of our assumptions that go into the software, as well as a first attempt to constrain some of the high level specification for a hypothetical but realistic multi-object integral field spectrograph assisted by MOAO system. Because 2D kinematics (i.e., velocity field and velocity dispersion maps) are one of the most obvious and easiest to currently simulate, they provide the most straight forward way of constraining the high level science design requirements for this type of instrument. We emphasize that the goal of this paper is not to study the detailed scientific capability of such instruments. Instead, the present paper aims at examining a few scientifically motivated cases to derive relevant specifications for such types of instruments with a capable MOAO system. These specifications will then be adopted as a baseline for studying their scientific capabilities subsequent papers. This paper is organized as follows: in Section 2, we detail the goals of this study and our methodology; in Sect. 3 we present the new simulation pipeline and in Sect. 4 how MOAO PSFs are simulated; In Sect. 5 we describe the kinematic measurements; In Sect. 6 and 7, we present the simulations and their results, which are discussed in Sect. 8. In Sect. 9, we draw the conclusions of this study. ", "conclusions": "We have developed software capable of end-to-end simulations of integral field spectrograph fed by an MOAO system. We have used this software to investigate a limited number of simulations in H-band, in order to give insights into how such MOAO-fed systems should perform on the VLT and the E-ELT. By requiring that the instrument is able to distinguish between a rotating disk and a major merger, we were able to constrain the required Ensquared Energy to be able to recover their large-scale motions. Separating the progenitors of the major merger requires only modest EE (15-26\\% at z=4 in a sample of 100-150mas; 30\\% at z=1.6 in 0.25 arcsec), but long exposure times ($\\sim$24hr). Higher EEs are generally needed to recover the full 2D kinematics in our simulations, typically 35\\%. Provided that the total SNR is large enough over the galaxy (typically $\\sim$ 5), it is possible to use more sophisticated methods, such as the diagnostic diagram introduced here, to distinguish between both types of objects, even at relatively low integration time for such distant objects (8 hrs). To generalize these results beyond what we described here, we have related these specifications to the concept of ``PSF contrast'', which is the amount of polluting light from adjacent spectra and which drives the minimal EE at a given spatial scale. The choice of the latter is driven mostly by the ``scale-coupling'' which is the relative size of the IFU pixel and the size of the features that the observe wishes to recovered in the data cube. Distinguishing between a major merger and a disk, for example, can be largely done by investigating the large scale motions within a system, and a relatively coarse spatial resolution appears then to be sufficient. Given this situation, a pixel scale of 50-75 mas seems to be a relatively good choice, since it provides sufficient SNR in less time than a finer sampling and also has the advantage of relaxing the MOAO system requirements. More stringent requirements can be set by attempting to resolve and investigate structure in high redshift galaxies and that will be simulated in subsequent papers." }, "0808/0808.0755.txt": { "abstract": "We present a spectroscopic analysis of five stellar streams (`A', `B', `Cr', `Cp' and `D') as well as the extended star cluster, \\ec4, which lies within \\streamc, all discovered in the halo of M31 from our CFHT/MegaCam survey. These spectroscopic results were initially serendipitous, making use of our existing observations from the DEep Imaging Multi-Object Spectrograph mounted on the Keck~II telescope, and thereby emphasizing the ubiquity of tidal streams that account for $\\sim$70\\% of the M31 halo stars in the targeted fields. Subsequent spectroscopy was then procured in \\streamc\\ and \\streamd\\ to trace the velocity gradient along the streams. Nine metal-rich ([Fe/H]$\\sim$-0.7) stars at $v_{\\rm hel}=-349.5$~km/s, $\\sigma_{v, corr}\\sim5.1\\pm2.5$\\,km/s are proposed as a serendipitous detection of \\streamcr, with followup kinematic identification at a further point along the stream. Six metal-poor ([Fe/H]$\\sim$-1.3) stars confined to a narrow, 15~km/s velocity bin centered at $v_{\\rm hel}=-285.6$~km/s, $\\sigma_{v,corr}=4.3^{+1.7}_{-1.4}$~km/s represent a kinematic detection of \\streamcp, again with followup kinematic identification further along the stream. For the cluster \\ec4, candidate member stars %consistent with the HST-constrained red giant branch with average [Fe/H]$\\sim$-1.4 ([Fe/H]$_{spec}$=-1.6), are found at $v_{\\rm hel}=-285$~km/s suggesting it could be related to \\streamcp. No similarly obvious cold kinematic candidate is found for \\streamd, although candidates are proposed in both of two spectroscopic pointings along the stream (both at $\\sim-400$km/s). Spectroscopy near the edge of \\streamb\\ suggests a likely kinematic detection at $v_{\\rm hel}\\sim-330$~km/s, $\\sigma_{v, corr}\\sim6.9$\\,km/s, while a candidate kinematic detection of \\streama\\ is found (plausibly associated to M33 rather than M31) with $v_{\\rm hel}\\sim-170$~km/s, $\\sigma_{v, corr}=12.5$\\,km/s. The low dispersion of the streams in kinematics, physical thickness, and metallicity makes it hard to reconcile with a scenario whereby these stream structures as an ensemble are related to the giant southern stream. We conclude that the M31 stellar halo is largely made up of multiple kinematically cold streams. ", "introduction": "\\footnotetext[1]{The data presented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W.M. Keck Foundation.} %-giant streams and dwarfs intro %{\\bf PERHAPS still to add: mass measurements of the streams; chemical analysis of the stream stars on average relative to the background halo population in these fields - no differences photometrically? what about spectral features?} Stellar streams represent the visible debris of small galaxies being cannibalized by large galaxies, memorials to the merging process by which the halos of galaxies are built up. The best known examples of streams in the Milky Way (MW) have recently been mapped far more extensively by the SDSS-DR5 by Belokurov et al.\\ (2006, 2007): the tidally stripped stars and globular clusters associated with the Sagittarius dwarf spheroidal and the Low Latitude stream, along with a newly discovered ``Orphan Stream'' so named for its lack of obvious progenitor. In M31, thanks to our ability to efficiently map vast regions of the halo, the number of discovered giant streams already outnumbers that of the MW (Ibata et al.\\ 2007). However, of the eight stellar streams that have been identified in the halo of M31, only one has plausibly been identified to a dwarf satellite: the loop connecting to NGC~205 (McConnachie et al.\\ 2005). This suggests that the other streams could represent an additional seven `uncataloged' satellites, although some of the streams might be produced by a common progenitor, as suggested by models of Fardal et al.\\ (2007, 2008) for the M31 Giant Southern Stream, or in a similar fashion to the Sgr dwarf and its numerous wraps around the Milky Way. It is important to characterize their orbits, metallicities and masses, to understand what their progenitors must have been. % These streams are still recognizable as ... The existence of stellar streams tells us that a progenitor galaxy has undergone significant mass loss. This is due to a combination of its orbit and its phase of evolution -- the amount of dark matter mass the galaxy has lost (Pe\\~narrubia et al.\\ 2008a,b). % %Dwarfs which are on more circular orbits are in principle less likely to %be in a phase of catastrophic mass loss (although interactions with the disk can lead to their disruption -- Pe\\~narrubia et al.\\ 2006), but Satellites on circular orbits are harder to disrupt, but if they are massive enough (i.e.\\ of the order of the LMC), dynamical friction will bring them close to the host galaxy centre, where the interactions with the disc will lead to their tidal disruption -- e.g., Pe\\~narrubia et al.\\ 2007). In addition, to form a stream one has to remove most of the dark matter halo ($\\sim$90--99\\%). The most important parameter that controls the mass loss rate of a dSph is the pericentre distance (the orbital eccentricity is of second order). On the other hand dwarfs with highly elliptical orbits spend a lot of time near apocentre where they are unlikely to be disrupted by the host. It is therefore not immediately obvious that streams represent preferred types of orbits on average. However, a number of theoretical studies have shown that significant information about the orbital properties of the progenitor galaxy can be derived from the streams (e.g, Ghigna et al.\\ 1998; Helmi et al.\\ 1999a). Streams can be much more informative to study than dwarfs because their orbits can be directly traced and constrained. Fellhauer et al.\\ (2006) were able to accurately constrain the shape of the Galactic potential through the bifurcation of Sagittarius streams in Belokurov et al.\\ (2006). Understanding the range of orbits of satellites to large galaxies will help us to understand how the halos of these galaxies formed. This is especially interesting in light of M31's huge stellar halo reflecting the dark matter dominated halo out to $\\simgt$150~kpc (e.g., Irwin et al.\\ 2005, Gilbert et al.\\ 2006, Ibata et al.\\ 2007). However streams can also be much harder to analyse observationally: the distances are problematic, there's a much lower spatial density and they have a larger extent so that observational sampling is not trivial. There is also the difficulty to infer the membership of different stream pieces, especially if we expect different chemical signatures due to metallicity gradients in the progenitor system (e.g., Ibata et al.\\ 2007). %M31 has 16 dwarf galaxy satellites, along with M33 which is likely %orbiting M31 as well. %Importantly, the faint dSph satellites of the MW and M31 begin %to depart significantly from the established correlations between %mass and metallicity, and M/L versus light. Streamy/blobby structures %(other than the giant stream) are individually interesting and constraining for the halo formation. They can represent the only traceable product of long disintegrated progenitors, yet retain a coherent body for statistical analysis. Streams can provide important clues on the structure of the progenitors (e.g.\\ metallicity gradients, mass to light -- M/L) as well as on the shape of the host dark matter halo (e.g.\\ prolate versus oblate) (Martinez-Delgado et al.\\ 2008). Future study of these structures will be able to put them in a much better \"near-field cosmology\" context, eventually understanding their ages and chemical histories. However it is important to uncover and study them now, even in limited capacities necessitated by the small numbers of spectroscopically identifiable stars and HR-diagram depths, so we can build our models on the most complete context. We have initiated a spectroscopic survey of the new streams found in M31's halo using the DEep Imaging Multi-Object Spectrograph on Keck~II to derive radial velocities and metallicities of red giant branch (RGB) stars. In this contribution, we discuss spectroscopic pointings in each of streams `C' and `D' which we obtained by serendipity, since the spectroscopy was taken prior to knowledge of the photometrically discovered streams, as well as followup spectroscopic pointings in both of these streams. We also analyze spectroscopic data from Koch et al.\\ (2008) lying within the Ibata et al.\\ (2007) streams `A' and `B'. %{\\bf %shows how streams really must be everywhere? quantify how lucky we were?} %summarize the characteristics of the stream in the %intro (from Ibata et al. 07) %-extent, metallicities etc. ", "conclusions": "In conclusion, we have conducted a Keck/DEIMOS spectroscopic survey of five stellar streams, recently uncovered through deep imaging observations of the halo. $\\bullet$ We have uncovered a kinematic substructure at \\vhel=-349.5$\\pm1.8$~km/s from a spectroscopic field lying in the Ibata et al.\\ (2007) \\streamc. The cold component has $\\sigma_{v_r}=5.1\\pm2.5$ and a narrow range in [Fe/H]=-0.7$\\pm$0.2, which we propose represents a metal-rich component, \\streamcr. $\\bullet$ We have uncovered a second kinematic substructure in the same field as \\streamcr\\ at \\vhel=-285.6$\\pm1.2$~km/s with $\\sigma_{v_r}=4.3^{+1.7}_{-1.4}$~km/s (non-zero at $>$3$\\sigma$ confidence interval) and a narrow range in [Fe/H]=-1.3$\\pm$0.2, which we propose represents a metal-poor stream, \\streamcp. We demonstrated that this kinematic \\streamcp\\ has a counterpart in a spatially offset metal-poor region of \\streamc\\ in Ibata et al.\\ (2007). $\\bullet$ We plausibly detect both \\streamcr\\ and \\streamcp\\ at a position $\\sim$30kpc further north along the structure, with no detectable velocity gradiant for \\streamcr, and a measured velocity gradient of $\\sim$40~km/s for \\streamcp. %the majority of stars even in 305TaS must give rise to the photometric contrast, and thus % it is likely the stars of similar Fe/H in the -400km//s range are the continuation of this stream. $\\bullet$ We were unable to identify kinematic substructure unambiguously associated to \\streamd\\ from our serendipitous spectroscopic pointing, however subsequent spectroscopy well centered in the \\streamd\\ identifies a likely cold kinematic structure which has a viable counterpart in the serendipitous pointing. We propose a kinematic detection of \\streamd\\ at \\vhel=-390.5~km/s with $\\sigma_{v_r}=4.2$~km/s. $\\bullet$ Spectroscopy near the edges of \\streama\\ and \\streamb\\ suggest a likely kinematic detection for \\streamb\\ with $v_{\\rm hel}\\sim-330$~km/s, $\\sigma_{v, corr}\\sim6.9$\\,km/s, and a kinematic detection of \\streama\\ at $v_{\\rm hel}\\sim-172$~km/s, $\\sigma_{v_r}\\sim12.5$\\,km/s. Neither spectroscopic pointing in these streams is ideally placed, and additional spectroscopic observations are well motivated to further constrain the kinematics of these structures. $\\bullet$ The extended cluster EC4 lies in the \\streamc\\ region, with kinematics (v$_{\\rm hel}$=-285\\,km/s) %, $\\sigma_{v, corr}\\sim3$~km/s), and metallicity ([Fe/H]=-1.4) which suggest it is related to the more metal-poor stream \\streamcp. %The resolved velocity dispersion of \\ec4\\ suggests the possible presence of a sizable dark matter component (Collins et al.\\ 2008a). %with M/L$\\sim$35, although the results are clearly preliminary based on only 8 member stars. \\ec4\\ could be the progenitor of the metal-poor \\streamcp\\ (somewhat unlikely given the apparent stellar mass difference between the stream and \\ec4), or it may simply be a structure carried along by the disrupted stream progenitor. In this case, and if \\ec4\\ has a sizable dark matter component, we have in fact detected the very first {\\it sub-sub-halo} (i.e.\\ a galaxy that was bound to a satellite galaxy), possibly explaining its small (r$_c$=30~pc) size. $\\bullet$ By explicitly removing stars belonging to the streams by their kinematics we can assess the underlying M31 stellar halo density and metallicity on the minor axis. This contrasts the purely photometric approach where Galactic contamination and stars belonging to stream substructures can only be removed statistically. Our resulting halo measurement has so few stars as to be highly uncertain statistically, however, it does reveal the general power of kinematic analysis of the halo population for future endeavors. The fraction of background halo stars in these stream fields suggests the conclusion that stellar halos are largely made up of multiple kinematically cold streams." }, "0808/0808.1369_arXiv.txt": { "abstract": "We suggest a scenario where the three light quark flavors are sequentially deconfined under increasing pressure in cold asymmetric nuclear matter as\\eg in neutron stars. The basis for our analysis is a chiral quark matter model of Nambu--Jona-Lasinio (NJL) type with diquark pairing in the spin-1 single flavor (CSL), spin-0 two flavor (2SC) and three flavor (CFL) channels. We find that nucleon dissociation sets in at about the saturation density, $n_0$, when the down-quark Fermi sea is populated (d-quark dripline) due to the flavor asymmetry induced by $\\beta$-equilibrium and charge neutrality. At about $3n_0$ u-quarks appear and a two-flavor color superconducting (2SC) phase is formed. The s-quark Fermi sea is populated only at still higher baryon density, when the quark chemical potential is of the order of the dynamically generated strange quark mass. We construct two different hybrid equations of state (EoS) using the Dirac-Brueckner Hartree-Fock (DBHF) approach and the EoS by Shen \\etal in the nuclear matter sector. The corresponding hybrid star sequences have maximum masses of, respectively, 2.1 and 2.0 M$_\\odot$. Two- and three-flavor quark-matter phases exist only in gravitationally unstable hybrid star solutions in the DBHF case, while the Shen-based EoS produce stable configurations with a 2SC phase component in the core of massive stars. Nucleon dissociation due to d-quark drip at the crust-core boundary fulfills basic criteria for a deep crustal heating process which is required to explain superbusts as well as cooling of X-ray transients. ", "introduction": "The phenomenology of compact stars is intimately connected to the EoS of matter at densities well beyond the nuclear saturation density, $n_0=0.16$ fm$^{-3}$. Compact stars are therefore natural laboratories for the exploration of baryonic matter under extreme conditions, complementary to those created in terrestrial experiments with atomic nuclei and heavy-ion collisions. Recent results derived from observations of compact stars provide serious constraints on the nuclear EoS, see \\cite{Klahn:2006ir} and references therein. A stiff EoS at high density is needed to explain the high compact-star masses, $M\\sim 2.0$~M$_\\odot$, reported for some low-mass X-ray binaries (LMXBs)\\eg 4U 1636-536 \\cite{Barret:2005wd}, and the large radius, $R > 12$~km, of the isolated neutron star RX J1856.5-3754 (shorthand: RX J1856) \\cite{Trumper:2003we}. Another example is EXO 0748-676, an LMXB for which the compact-star mass {\\it and} radius have been constrained to $M\\ge 2.10\\pm 0.28$~M$_\\odot$ and $R \\ge 13.8 \\pm 0.18$~km \\cite{Ozel:2006km}. However, the status of the results for the latter object is unclear, because the gravitational redshift $z=0.35$ observed in the X-ray burst spectra \\cite{Cottam:2002} has not been confirmed, despite numerous attempts. While compact-star phenomenology apparently points towards a stiff EoS at high density, heavy-ion collision data for kaon production \\cite{Fuchs:2005zg} and elliptic flow \\cite{Danielewicz:2002pu} set an upper limit on the stiffness of the EoS \\cite{Klahn:2006ir}. A key question regarding the structure of matter at high density is whether a phase transition to quark matter occurs inside compact stars, and whether it is accompanied by unambiguous observable signatures. It has been argued that the observation of a compact star with high mass and large radius, likewise reported for EXO 0748-676, would be incompatible with a quark core \\cite{Ozel:2006km}, because the softening of the EoS due to quark deconfinement lowers the maximum mass and the radius in comparison to the nuclear matter case. However, Alford \\etal \\cite{Alford:2006vz} have demonstrated with a few counter examples that quark matter cannot be excluded by this argument. In particular, for a recently developed hybrid star EoS \\cite{Klahn:2006iw}, based on the DBHF approach in the nuclear sector and a three-flavor chiral quark model \\cite{Blaschke:2005uj}, stable hybrid stars in the mass range from 1.2 M$_\\odot$ up to 2.1 M$_\\odot$ have been obtained, in accordance with modern mass-radius constraints, see also \\cite{Blaschke:2007ri}. In that model, a sufficiently low critical density for quark deconfinement was achieved with a strong diquark coupling, while a sufficient stiffness for a high maximum mass of the compact star sequence was obtained with a repulsive vector meanfield in the quark matter sector. The corresponding hybrid EoS for symmetric matter was shown to fulfill the constraints from elliptic flow data in heavy-ion collisions. In the present work we want to discuss a new scenario of quark deconfinement, which could play an important role in asymmetric matter, in particular for the phenomenology of compact stars. Chiral quark models of the NJL type with dynamical chiral symmetry breaking have the property that the restoration of this symmetry (and the related quark deconfinement) at zero temperature is flavor specific. When solving the gap and charge-neutrality equations selfconsistently one finds that the chiral symmetry restoration for a given flavor occurs when the chemical potential of that flavor reaches a critical value that is approximately equal to the dynamically generated quark mass, $\\mu_{f}=\\mu_c \\approx m_{f}$, where $f=u, d, s$. In asymmetric matter the quark chemical potentials are different. Consequently, the NJL model behavior suggests that the critical density of deconfinement is flavor dependent, see Fig. \\ref{f:phases}. \\begin{figure} % \\begin{tabular}{ll} \\includegraphics[angle=0,width=0.5\\textwidth]{d-drip2.eps}& \\includegraphics[angle=0,width=0.4\\textwidth,clip=]{CSL-gaps.eps} \\end{tabular} \\caption{Left panel: Chemical potentials of up and down quarks (strange quark sector not shown). With increasing quark chemical potential $\\mu_q=(\\mu_u+\\mu_d)/2$ in isospin asymmetric matter the quark flavors pass sequentially the threshold ($\\mu_c$) for chiral symmetry restoration (deconfinement), which entails nucleon dissociation. Right panel: Solution of the NJL gap equations under isospin asymmetry.} \\label{f:phases} \\end{figure} In this approach the down quark flavor is the first to drip out of nucleons as the density increases, followed by the up quark flavor and eventually also by strange quarks. This behavior is absent in simple and commonly applied thermodynamic bag model equations of state since they are essentially flavor blind. Under the $\\beta$-equilibrium condition in compact stars the chemical potentials of quarks and electrons are related by $\\mu_d=\\mu_s$ and $\\mu_d=\\mu_u+\\mu_e$. The mass difference between the strange and the light quark flavors $m_s \\gg m_u, m_d$ has two consequences: (1) the down and strange quark densities are different, so charge neutrality requires a finite electron density and, consequently, (2) $\\mu_d>\\mu_u$. When increasing the baryochemical potential, the d-quark chemical potential is therefore the first to reach the critical value $\\mu_c$ where the chiral symmetry gets (approximately) restored in a first-order transition and a finite density of d-quarks appears. Due to the finite value of $\\mu_e$, the u-quark chemical potential is still below $\\mu_c$ while the s-quark density is zero due to the high s-quark mass. A {\\it single-flavor} d-quark phase therefore forms in co-existence with the positively charged nuclear-matter medium. Why has this interesting scenario been left unnoticed? On the one hand, bag models, which are commonly used to describe quark matter in compact star interiors cannot address sequential deconfinement. On the other hand, the single-flavor d-quark phase is negatively charged and cannot be neutralized in a purely leptonic background. This was a reason to disregard it in dynamical approaches like the NJL models. In the following we discuss the single-flavor phase for the first time under the natural assumption that the neutralizing background is nuclear matter. Since nucleons are bound states of quarks, the physical context in which such a mixed phase of nucleons and free d-quarks occurs is that of the dissociation of nucleonic bound states of quarks (Mott effect). ", "conclusions": "In this contribution we have suggested a new quark-nuclear hybrid EoS for compact star applications that fulfills modern observational constraints from compact stars. Due to isospin asymmetry, down-quarks may ``drip out'' from nucleons and form a single-flavor color superconducting (CSL) phase that is mixed with nuclear matter already at the crust-core boundary in compact stars. The CSL phase has interesting cooling and transport properties that are in accordance with constraints from the thermal and rotational evolution of compact stars. It remains to be investigated whether this new compact star composition could lead to unambiguous observational consequences. We conjecture that the d-quark drip may serve as an effective deep crustal heating mechanism for the explanation of the puzzling superburst phenomenon and the cooling of X-ray transients." }, "0808/0808.3464_arXiv.txt": { "abstract": "{ We present the AGILE gamma-ray observations of the field containing the puzzling gamma-ray source 3EG J1835+5918. This source is one of the most remarkable unidentified EGRET sources. } { An unprecedentedly long AGILE monitoring of this source yields important information on the positional error box, flux evolution, and spectrum. }{3EG J1835+5918 has been in the AGILE field of view several times in 2007 and 2008 for a total observing time of 138 days from 2007 Sept 04 to 2008 June 30 encompassing several weeks of continuous coverage.}{With an exposure time approximately twice that of EGRET, AGILE confirms the existence of a prominent gamma-ray source (\\sagile) at a position consistent with that of EGRET, although with a remarkably lower average flux value for photon energies greater than 100 MeV. {A 5-day bin temporal analysis of the whole data set of \\sagile~ shows some evidence for variability of the gamma-ray flux. } The source spectrum between 100 MeV and 1 GeV can be fitted with a power law with photon index in the range 1.6-1.7, fully consistent with the EGRET value.} { The faint X-ray source RX J1836.2+5925 that has been proposed as a possible counterpart of \\s$\\;$ is well within the AGILE error box. Future continuous monitoring (both by AGILE and GLAST) is needed to confirm the gamma-ray flux variability and to unveil the source origin, a subject that is currently being pursued through a multiwavelength search for counterparts. } ", "introduction": " ", "conclusions": "The AGILE observations and monitoring of \\sagile$\\;$ adds relevant information about this puzzling source. Postponing an account of our multifrequency observations of the region to a forthcoming paper, we briefly outline here a few important points regarding the search for a counterpart. We confirm a point already noticed by other authors, i.e., the absence of a blazar or any other relatively bright radio sources in the field containing the AGILE (and EGRET) error box. The analysis of Mattox et al. (2001) shows a 100 mJy source (B1834+5904) positioned at the 99.5$\\%$ probability contour ($12'.7$ from the EGRET centroid position and $21'.8$ from the AGILE centroid position). By using the NED, SIMBAD, and the Massaro et al. (2007) catalogues, we confirm the absence of a radio-loud blazar in the revised AGILE error box. Outside this error box, we notice the existence of RGB J1841+591, a BL Lac type object positioned at $43'.5$ from the AGILE centroid position. Even though occasional contributions from this object cannot be excluded in our analysis of \\agile, we emphasise that the AGILE integrated flux positioning at the 95\\% contour level is clearly not consistent with any substantial contribution from RGB~J1841+591. We also notice that the faint X-ray source RX J1836.2+5925 that was proposed as a possible counterpart is well within the AGILE error box of \\sagile. A hint of variability of this source was noticed when comparing two HRI ROSAT observations taken almost three years apart \\citep{Mirabal_2001}, prompting a claim for the discovery of a new class of compact gamma-ray sources \\citep{Mirabal_2000}. However, subsequent Chandra X-ray observations showed a practically constant X-ray flux of RX J1836.2+5925, and weakened the variability claim in favour of a scenario encompassing a constant Geminga-like source. We note here that our group observed the region containing RX J1836.2+5925 with the SWIFT XRT and XMM-Newton during the period May-June 2008. Analysis of these data and a full discussion of the \\sagile$\\;$ counterpart problem will be presented elsewhere. Future continuous monitoring (both by AGILE and GLAST) is needed to confirm the gamma-ray flux variability and to unveil the source origin." }, "0808/0808.1461_arXiv.txt": { "abstract": "Magnetic fields in the early universe can significantly alter the thermal evolution and the ionization history during the dark ages. This is reflected in the $21$ cm line of atomic hydrogen, which is coupled to the gas temperature through collisions at high redshifts, and through the Wouthuysen-Field effect at low redshifts. We present a semi-analytic model for star formation and the build-up of a Lyman $\\alpha$ background in the presence of magnetic fields, and calculate the evolution of the mean $21$ cm brightness temperature and its frequency gradient as a function of redshift. We further discuss the evolution of linear fluctuations in temperature and ionization in the presence of magnetic fields and calculate the effect on the $21$ cm power spectrum. At high redshifts, the signal is increased compared to the non-magnetic case due to the additional heat input into the IGM from ambipolar diffusion and the decay of MHD turbulence. At lower redshifts, the formation of luminous objects and the build-up of a Lyman $\\alpha$ background can be delayed by a redshift interval of $10$ due to the strong increase of the filtering mass scale in the presence of magnetic fields. This tends to decrease the $21$ cm signal compared to the zero-field case. In summary, we find that $21$ cm observations may become a promising tool to constrain primordial magnetic fields. ", "introduction": "{Observations of the $21$ cm fine structure line of atomic hydrogen have the potential to become an important means of studying} the universe at early times, during and even before the epoch of reionization. This possibility was suggested originally by \\citet{Purcell}, and significant process in instrumentation and the development of radio telescopes has brought us close to the first observations from radio telescopes like LOFAR\\footnote{http://www.lofar.org}. While one of the main purposes is to increase our understanding of cosmological reionization \\citep{Bruyn}, a number of further exciting applications have been suggested in the mean time. \\citet{Loeb} demonstrated how $21$ cm measurements can probe the thermal evolution of the IGM at a much earlier time, at redshifts of $z\\sim200$. \\citet{Barkana} suggested a method that separates physical and astrophysical effects and thus allows to probe the physics of the early universe. \\citet{FurlanettoDM} showed how the effect of dark matter annihilation and decay would be reflected in the $21$ cm line, and effects of primordial magnetic fields have been considered by \\citet{Tashiro}.\\\\ \\\\ Indeed, primordial magnetic fields can affect the early universe in various ways. The thermal evolution is significantly altered by ambipolar diffusion heating and decaying MHD turbulence \\citep{Sethi05, Sethi08, Schleicher}. \\citet{Kim} calculated the effect of the Lorentz force on structure formation and showed that additional power is present on small scales in the presence of primordial magnetic fields. It was thus suggested that reionization occurs earlier in the presence of primordial magnetic fields \\citep{Sethi05, TashiroReion}. However, as pointed out by \\citet{Gnedin}, the characteristic mass scale of star forming halos, the so-called filtering mass, increases signficantly when the temperature is increased. For comoving field strengths of $\\sim1$ nG, we found that the filtering mass scale is shifted to scales where the power spectrum is essentially independent of the magnetic field \\citep{Schleicher}. Reionization is thus delayed in the presence of primordial magnetic fields. We further found upper limits of the order $1$ nG, based on the Thomson scattering optical depth measured by WMAP 5 \\citep{Komatsu, WMAPAngular} and the requirement that reionization ends at $z\\sim6$ \\citep{Becker}. The presence of primordial magnetic fields can have interesting implications on first star formation as well. In the absence of magnetic fields, \\citet{Abel} and \\citet{Bromm} suggested that the first stars should be very massive, perhaps with $\\sim100$ solar masses. \\citet{Clark} and \\citet{Omukai} argued that in more massive and perhaps metal-enriched galaxies, fragmentation should be more effective and lead to the formation of rather low-mass stars, due to a stage of efficient cooling \\citep{Omukai}. {Magnetic fields may change this picture and reduce the stellar mass by triggering jets and outflows \\citep{Silk}. However, simulations by \\citep{Machida} show that the change in mass is of the order $10\\%$.} \\\\ \\\\ $21$ cm measurements can try to adress {primordial magnetic fields} in two ways: During the dark ages of the universe, at redshifts $z\\sim200$ well before the formation of the first stars, the spin temperature of hydrogen is coupled to the gas temperature via collisional de-excitation by hydrogen atoms \\citep{Allison, Zygelman} and free electrons \\citep{Smith, FF07}, constituting a probe at very early times. While collisional de-excitation becomes inefficient due to the expansion of the universe, the first stars will build up a Lyman $\\alpha$ background that will cause a deviation of the spin temperature from the radiation temperature by the Wouthuysen-Field effect \\citep{Wouthuysen, Field}. As primordial magnetic fields may shift the onset of reionization, the onset of this coupling constitutes an important probe on the presence of such fields. We adress these possibilities in the following way: In \\S\\ref{IGM}, we review our treatment of the IGM in the presence of primordial magnetic fields. The evolution of the $21$ cm background and the role of Lyman $\\alpha$ photons is discussed in \\S\\ref{21cm}. The evolution of linear perturbations in temperature and ionization is calculated in \\ref{fluctuations}. The results for the power spectrum are given in \\S\\ref{power}, and the results are further discussed in \\S\\ref{discussion}. ", "conclusions": "\\label{discussion} In the previous sections, we have presented a semi-analytic model describing the post-reionization universe and reionization in the context of primordial magnetic fields, and calculated the consequences for the mean $21$ cm brightness fluctuation and the large-scale power spectrum. formation of We identify two regimes in which primordial magnetic fields can influence effects measured with $21$ cm telescopes. At low redshifts, primordial magnetic fields tend to delay reionization and the build-up of a Lyman $\\alpha$ background, thus shifting the point where the signal is at its maximum, and changing the amplitude of the $21$ cm power spectrum. The first $21$ cm telescopes like LOFAR and others will focus mainly on the redshift of reionization and can thus probe the epoch when a significant Lyman $\\alpha$ background builds up. As our understanding of the first stars increases due to advances in theoretical modeling or due to better observational constraints, this may allow us to determine whether primordial magnetic fields are needed to delay the build-up of Lyman $\\alpha$ photons or not. As demonstrated above, comoving field strengths of the order $1$ nG can delay the build-up of a Lyman $\\alpha$ background by $\\Delta z\\sim 10$, which is significantly stronger than other mechanisms that might delay the formation of luminous objects. Lyman Werner feedback is essentially self-regulated and never leads to a significant suppression of star formation \\citep{Johnson, JohnsonO}, X-ray feedback from miniquasars is strongly constrained from the observed soft X-ray background \\citep{Salvaterra}, and significant heating from dark matter decay or annihilation would be accompanied by a significant amount of secondary ionization, resulting in a too large optical depth. On the other hand, an important issue is the question regarding the first sources of light. As shown in \\citet{Schleicher}, massive Pop. III stars are needed to provide the correct reionization optical depth. On the other hand, an additional population of less massive stars with an IMF according to \\citep{Scalo,Kroupa,Chabrier} might be present. Such a population would emit more photons between the Lyman $\\alpha$ and $\\beta$ line per stellar baryon \\citep{Leitherer}, and could thus shift the build-up of a Lyman $\\alpha$ background to an earlier epoch. The onset of efficient coupling via the Wouthuysen-Field effect thus translates into a combined constraint on the stellar population and the strength of primordial magnetic fields.\\\\ \\\\ A further effect occurs at high redshifts, where the additional heat input from magnetic fields due to ambipolar diffusion and the decay of MHD turbulence increases the $21$ cm signal. As gas decouples earlier from the radiation field, the difference between gas and radiation temperature is larger and collisions are more effective in coupling the spin temperature to the gas temperature. The determination of the $21$ cm signal from this epoch is certainly challenging, as the foreground emission corresponds to temperatures which are higher than the expected $21$ cm brightness temperature by several orders of magnitude. However, as pointed out in other works \\citep{DiMatteo, OhMack, Zaldarriaga, Sethi, Yu}, the foreground emission is expected to be featureless in frequency, which may allow for a sufficiently accurate subtraction. In this context, it may also to help to focus on the frequency gradient of the mean brightness temperature, rather than the $21$ cm brightness temperature itself. Upcoming long-wavelength experiments such as LOFAR, 21CMA (former PAST)\\footnote{http://web.phys.cmu.edu/~past}, MWA\\footnote{http://web.haystack.mit.edu/arrays/MWA/index.html}, LWA\\footnote{http://lwa.unm.edu} and SKA\\footnote{http://www.skatelescope.org} may thus detect the additional heat from primordial magnetic fields in the neutral gas, or otherwise set new upper limits on primordial magnetic fields, perhaps down to $B_0\\sim0.1$ nG. Like the 21 cm transition of hydrogen, rotational and ro-vibrational transitions of primordial molecules may create interesting signatures in the CMB as well\\citep{Schleicher2}, which may provide a further test of the thermal evolution during the dark ages." }, "0808/0808.1182_arXiv.txt": { "abstract": "We introduce the {\\it SMC in space and time}, a large coordinated space and ground-based program to study star formation processes and history, as well as variable stars, structure, kinematics and chemical evolution of the whole SMC. Here, we present the Colour-Magnitude Diagrams(CMDs) resulting from HST/ACS photometry, aimed at deriving the star formation history (SFH) in six fields of the SMC. The fields are located in the central regions, in the stellar halo, and in the wing toward the LMC. The CMDs are very deep, well beyond the oldest Main Sequence Turn-Off, and will allow us to derive the SFH over the entire Hubble time. ", "introduction": "The Small Magellanic Cloud (SMC) is the closest late-type dwarf and has many properties similar to those of the vast majority of this common class of galaxies. Its current metallicity (Z$\\simeq$0.004 in mass fraction, as derived from HII regions and young stars) is typical of dwarf irregular and Blue Compact Dwarf (BCD) galaxies, the least evolved systems, hence the most similar to primeval galaxies. Its mass (between 1 and 5 $\\times 10^9 M_{\\odot}$, e.g. \\cite[Kallivayalil et al. 2006]{K06} and references therein) is at the upper limit of the range of masses typical of late-type dwarfs. These characteristics, combined with its proximity, make the SMC the natural benchmark to study the evolution of late-type dwarf galaxies. Moreover, its membership to a triple system allows detailed studies of interaction-driven modulations of the star formation activity. A wealth of data on the SMC are available in the literature, although not as much as for its bigger companion, the LMC. Yet, much more are needed for a better understanding of how the SMC has formed and evolved. We have thus embarked on a long-term project to study the evolution of the SMC in space and time. Our project plans to exploit the high performances in depth, resolving power or large field of view of current and forthcoming, space and ground based, telescopes, such as HST, VLT, Spitzer, SALT and VST. ", "conclusions": "" }, "0808/0808.2462_arXiv.txt": { "abstract": "{ In supersymmetric models with a long-lived stau being the lightest Standard Model superpartner, the stau abundance during primordial nucleosynthesis is tightly constrained. Considering the complete set of stau annihilation channels in the minimal supersymmetric Standard Model (MSSM) with real parameters for scenarios in which sparticle coannihilations are negligible, we calculate the decoupling of the lighter stau from the primordial plasma and identify processes which are capable to deplete the resulting stau abundance significantly. We find particularly efficient stau annihilation at the resonance of the heavy CP-even Higgs boson and for a lighter stau with a sizeable left--right mixing due to enhanced stau-Higgs couplings. Even within the constrained MSSM, we encounter both effects leading to exceptionally small values of the resulting stau abundance. Prospects for collider phenomenology are discussed and possible implications of our findings are addressed with emphasis on gravitino dark matter scenarios. } \\begin{document} ", "introduction": "\\label{sec:introduction} The appearance of the lighter stau $\\stauone$ as the lightest Standard Model superpartner---or lightest ordinary superpartner (LOSP)---is a commonplace occurrence even in supersymmetric (SUSY) models with restrictive assumptions on the SUSY breaking sector such as the constrained minimal supersymmetric Standard Model (CMSSM). If the lightest supersymmetric particle (LSP) is assumed to be the LOSP, this parameter region is not considered because of severe upper limits on the abundance of massive stable charged particles~\\cite{Yao:2006px}. However, for example, in axino/gravitino LSP scenarios~\\cite{Pagels:1981ke,Borgani:1996ag,Bonometto:1989vh,Covi:1999ty,Steffen:2007sp} and in scenarios with broken R parity~\\cite{Allanach:2003eb,Allanach:2006st,Allanach:2007vi,Takayama:2000uz,Buchmuller:2007ui},% \\footnote{In this work we assume that R-parity is conserved.} the $\\stau$ LOSP becomes unstable and thereby a viable option. Indeed, supersymmetric models with a long-lived $\\stauone$ LOSP are particularly promising for collider phenomenology~\\cite{Drees:1990yw,Nisati:1997gb,Feng:1997zr,Ambrosanio:2000ik,Buchmuller:2004rq,Hamaguchi:2004df,Feng:2004yi,Brandenburg:2005he,DeRoeck:2005bw,Martyn:2006as,Steffen:2006hw,Ellis:2006vu,Hamaguchi:2006vu}: Since the $\\stauone$ LOSP could escape the collider detector as a quasi-stable muon-like particle, it can be associated with signatures that are very different from the excess in missing energy expected in neutralino LSP scenarios. In the early Universe the negatively charged LOSP $\\stauone$'s and the associated positively charged anti-staus $\\stauone^*$'s were in thermal equilibrium for temperatures of $T>m_{\\stauone}/20\\gtrsim T_{\\freezeout}$. At $T_{\\freezeout}$, the annihilation rate of the (by then) non-relativistic $\\stauone$'s becomes smaller than the Hubble rate so that they decouple from the thermal plasma. Thus, for $T\\lesssim T_{\\freezeout}$, their yield $Y_{\\stau}\\equiv(n_{\\stauone}+n_{\\stauone^*})/s$ is given approximately by $\\Ystau\\approx Y^{\\equil}_{\\stau}(T_{\\freezeout})$, where $n_{\\stau}^{(\\equil)}\\equiv n_{\\stauone}^{(\\equil)}+ n_{\\stauone^*}^{(\\equil)}$ is the (equilibrium) number density of both $\\stauone$ and $\\stauone^*$ and $s=2\\pi^2\\,g_{*S}\\,T^3/45$ the entropy density with \\gstarS\\ effective degrees of freedom. This thermal relic abundance $\\Ystau$ is subject to cosmological constraints in SUSY scenarios with a long-lived $\\stauone$ LOSP: \\begin{itemize} \\item In axino/gravitino LSP scenarios, $\\Ystau$ governs the non-thermally produced (NTP) relic density of axino/gravitino dark matter that originates from $\\stauone$ decays~\\cite{Borgani:1996ag,Covi:1999ty} $\\Omega_{\\widetilde{\\mathrm{a}}/\\gravitino}^{\\NTP} h^2 = m_{\\widetilde{\\mathrm{a}}/\\gravitino}\\, \\Ystau\\, s(T_0) h^2 / \\rho_{\\mathrm{c}}$ where $m_{\\widetilde{\\mathrm{a}}/\\gravitino}$ denotes the axino/gravitino LSP mass and $\\rho_c/[s(T_0)h^2]=3.6\\times 10^{-9}\\,\\GeV$~\\cite{Yao:2006px}. Thus, the dark matter density $\\Omega_{\\mathrm{dm}}$ which limits $\\Omega_{\\widetilde{\\mathrm{a}}/\\gravitino}^{\\NTP}$ from above implies an upper limit on $\\Ystau$ for a given $m_{\\widetilde{\\mathrm{a}}/\\gravitino}$. This limit can become particularly restrictive in the case of additional sizeable contributions to $\\Omega_{\\mathrm{dm}}$ such as the ones from thermal axino/gravitino production $\\Omega_{\\widetilde{\\mathrm{a}}/\\gravitino}^{\\TP}$~\\cite{Bolz:2000fu,Brandenburg:2004du,Pradler:2006qh,Rychkov:2007uq}. For example, for $m_{\\widetilde{\\mathrm{a}}/\\gravitino}=50~\\GeV$ and $\\Omega_{\\widetilde{\\mathrm{a}}/\\gravitino}^{\\TP}=0.99\\,\\Omega_{\\mathrm{dm}}$ ($0.9\\,\\Omega_{\\mathrm{dm}}$), one finds $\\Ystau<10^{-13}$ ($10^{-12}$); cf.\\ Fig.~13 of Ref.~\\cite{Steffen:2006hw}. \\item For $\\stauone$ decays during/after big bang nucleosynthesis (BBN), the Standard Model particles emitted in addition to the axino/gravitino LSP can affect the abundances of the primordial light elements. This leads to upper limits on $\\xi_{\\mathrm{em}/\\mathrm{had}}\\equiv \\epsilon_{\\mathrm{em}/\\mathrm{had}}\\,\\Ystau$ that depend on the stau lifetime $\\tau_{\\stauone}$~\\cite{Cyburt:2002uv,Kawasaki:2004qu,Jedamzik:2006xz}. Here $\\epsilon_{\\mathrm{em}/\\mathrm{had}}$ denotes the (average) electromagnetic/hadronic energy emitted in a single $\\stauone$ decay, which can be calculated with particle physics methods for a given model. Accordingly, the BBN constraints on $\\xi_{\\mathrm{em}/\\mathrm{had}}$ can be translated into upper limits on $\\Ystau$; cf.\\ Fig.~12 of Ref.~\\cite{Steffen:2006hw} (and Figs.~14 and~15 of Ref.~\\cite{Kawasaki:2008qe}) for associated $\\Ystau$ limits in gravitino LSP scenarios, which can be as restrictive as $\\Ystau<10^{-14}$ ($10^{-15}$). \\item The mere presence of the negatively charged $\\stauone$'s at cosmic times of $t\\gtrsim 5\\times 10^3\\,\\seconds$ can lead to ($^4$He$\\,\\stauone$) and ($^8$Be$\\,\\stauone$) bound states and can thereby allow for catalyzed BBN (CBBN) of $^6$Li and $^9$Be to abundances far above the ones obtained in standard BBN (SBBN)~\\cite{Pospelov:2006sc,Bird:2007ge,Pospelov:2007js,Pospelov:2008ta}. Indeed, confronting the abundances obtained in CBBN with observationally inferred bounds on the primordial abundances of $^9$Be (and $^6$Li) imposes restrictive upper limits of $\\Ystau\\lesssim 2\\times 10^{-15}$ ($2\\times 10^{-15}$\\,--\\,$2\\times 10^{-16})$ for $\\tau_{\\stauone}\\gtrsim 10^5\\,\\seconds$; cf.\\ Fig.~5 in Ref.~\\cite{Pospelov:2008ta} for $n_{\\stauone}=n_{\\stauone^*}$. \\end{itemize} For example, in gravitino LSP scenarios with the $\\stauone$ LOSP being the next-to-lightest supersymmetric particle (NLSP) and conserved R-parity, the listed cosmological constraints have been confronted with representative values~\\cite{Asaka:2000zh,Fujii:2003nr}% \\begin{align} \\label{eq:yield-approx} \\Ystau \\simeq (0.4-1.5) \\times 10^{-13} \\left( \\frac{\\mstauone}{100~\\GeV} \\right) \\ , \\end{align} which are in good agreement with the curves in Fig.~1 of Ref.~\\cite{Asaka:2000zh} that have been obtained for the case of a purely `right-handed' $\\stauone\\simeq\\stauR$ NLSP and a bino-like lightest neutralino, $\\neutralino\\simeq\\Bino$, with a mass of $m_{\\Bino}=1.1\\,\\mstauone$. Thereby, it has been found that the (C)BBN constraints impose the limit $\\taustau\\lesssim 5\\times 10^3\\,\\seconds$~\\cite{Pospelov:2006sc,Pospelov:2008ta} with severe implications in the collider-friendly region of $\\mstauone < 1~\\TeV$: (i)~The $\\tau_{\\stauone}$ limit disfavors the kinematical determination of $\\mgravitino$~\\cite{Steffen:2006wx} and thereby both the determination of the Planck scale at colliders~\\cite{Buchmuller:2004rq} and the method proposed to probe the maximum reheating temperature $\\TR$ at colliders~\\cite{Pradler:2006qh}. (ii)~Within the CMSSM, the $\\tau_{\\stauone}$ limit implies an upper limit on the reheating temperature of $\\TR\\lesssim 10^7~\\GeV$~\\cite{Pradler:2006hh,Pradler:2007is,Pradler:2007ar} that disfavors the viability of thermal leptogenesis with hierarchical heavy Majorana neutrinos~\\cite{Fukugita:1986hr,Davidson:2002qv,Buchmuller:2004nz,Blanchet:2006be,Antusch:2006gy}. (iii)~The $\\tau_{\\stauone}$ limit can point to a CMSSM mass spectrum which will be difficult to probe at the Large Hadron Collider (LHC)~\\cite{Cyburt:2006uv,Pradler:2006hh,Pradler:2007is,Pradler:2007ar}. Indeed, the $\\tau_{\\stauone}$ limit can be relaxed only with a significant reduction of~(\\ref{eq:yield-approx}) which has been presented explicitly so far only for non-standard cosmological scenarios with a very low value of~$\\TR$~\\cite{Takayama:2007du} or with late-time entropy production after $\\stauone$ decoupling and before BBN~\\cite{Pradler:2006hh,Hamaguchi:2007mp}.% \\footnote{Note that some implications of the $\\tau_{\\stauone}$ limit can be evaded not only by relaxing it but also by respecting it, e.g., R-parity violation can lead to $\\tau_{\\stauone}<5\\times 10^3\\,\\seconds$~\\cite{Buchmuller:2007ui}.} In this work we calculate the decoupling of the lighter stau from the primordial plasma by taking into account the complete set of stau annihilation channels in the MSSM with real parameters for SUSY spectra for which sparticle coannihilation is negligible. Using our own code for the computation of the resulting thermal relic stau abundance $\\Ystau$, we examine explicitly (i)~the effect of left--right mixing of the lighter stau, (ii)~the effect of large stau--Higgs couplings, and (iii)~stau annihilation at the resonance of the heavy CP-even Higgs boson $H^0$. We consider both the ``phenomenological MSSM'' (pMSSM) (see, e.g.,~\\cite{Djouadi:2002ze}) in which the soft SUSY breaking parameters can be set at the weak scale, and the CMSSM, in which the gaugino masses, the scalar masses, and the trilinear scalar couplings are assumed to take on the respective universal values $\\monetwo$, $\\mzero$, and $A_0$ at the scale of grand unification $\\mgut\\simeq 2\\times 10^{16}\\,\\GeV$. Within the framework of the pMSSM, we show examples in which $\\Ystau$ can be well below $10^{-15}$. Even within the CMSSM, we encounter regions with exceptionally small values of $\\Ystau\\lesssim 2\\times 10^{-15}$. The implications of these findings are discussed for scenarios with the gravitino LSP and the stau NLSP. We also address the viability of a $\\stauone$--$\\stauone^*$ asymmetry. Remarkably, we find that key quantities for the significant $\\Ystau$ reduction could be probed at both the LHC and the International Linear Collider (ILC). A calculation of the thermal relic abundance of long-lived staus has also been part of a recent thorough study~\\cite{Berger:2008ti} which focusses on gauge interactions and on the effect of Sommerfeld enhancement. In contrast, the most striking findings of our study---in which Sommerfeld enhancement is not taken into account---are related to the Higgs sector of the MSSM. At this point, we should also stress that the \\texttt{micrOMEGAs} code~\\cite{Belanger:2001fz,Belanger:2004yn,Belanger:2007zz,Belanger:2008sj} allows for sophisticated calculations of the thermal relic stau abundance also in regions in which coannihilation effects become important. In fact, \\texttt{micrOMEGAs} has already been applied in several studies to calculate $\\Ystau$~\\cite{Fujii:2003nr,Pradler:2006hh,Kersten:2007ab,Pradler:2007ar,Berger:2008ti}. In this paper, we also work with \\texttt{micrOMEGAs} to cross check the results of our own $\\Ystau$ calculation and to calculate $\\Ystau$ in parameter regions in which sparticle coannihilations become relevant. The outline of this paper is as follows. In the next section we review basic properties of the staus to introduce our notations and conventions for the stau mixing angle. Section~\\ref{sec:prim-ann} explains the way in which we calculate $\\Ystau$ and provides the complete list of stau annihilation channels. In Sect.~\\ref{sec:dependence} we analyze the dependence of the most relevant stau annihilation channels on the stau mixing angle. Effects of large stau--Higgs couplings and stau annihilation at the $H^0$ resonance are studied in Sects.~\\ref{sec:enhanc-coupl-higgs} and~\\ref{sec:reson-annih}, respectively. The viability of a $\\stauone$--$\\stauone^*$ asymmetry is addressed in Sect.~\\ref{sec:comm-stau-stau}. In Sect.~\\ref{sec:annih-chann} we present exemplary parameter scans within the CMSSM that exhibit exceptionally small $\\Ystau$ values. Potential collider phenomenology of the parameter regions associated with those exceptional relic abundances and potential implications for gravitino dark matter scenarios are discussed in Sects.~\\ref{sec:collider} and~\\ref{sec:gravitino}, respectively. ", "conclusions": "\\label{sec:conclusions} Supersymmetric models with a long-lived stau $\\stauone$ being the lightest Standard Model superpartner are well-motivated and very attractive in light of potentially striking signatures at colliders. For a standard thermal history with primordial temperatures $T>\\mstauone/20>\\Tf$---which is the working hypothesis is this work---the long-lived $\\stauone$ becomes an electrically charged thermal relic whose abundance can be restricted by cosmological constraints. We have carried out a thorough study of primordial stau annihilation and the associated thermal freeze out. Taking into account the complete set of stau annihilation channels within the MSSM with real parameters for cases with negligible sparticle coannihilation, the resulting thermal relic $\\stauone$ yield $\\Ystau$ has been examined systematically. While related earlier studies focussed mainly on the $\\stauone\\simeq\\stauR$ case~\\cite{Asaka:2000zh,Fujii:2003nr,Pradler:2006hh,Berger:2008ti}, we have investigated cases in which $\\stauone$ contains a significant admixture of $\\stauL$ including the maximal mixing case and $\\stauone\\simeq\\stauL$. We find that the variation of the stau mixing angle $\\thetastau$ does affect the relative importance of the different annihilation channels significantly but not necessarily the resulting $\\Ystau$ value for relatively small values of $\\tanb$. By increasing $\\tanb$, however, we encounter a dramatic change of this picture for large absolute values of the Higgs-higgsino mass parameter $\\mu$ and/or of the trilinear coupling $\\Atau$, which are the dimensionful SUSY parameters that govern simultaneously stau left-right mixing and the stau--Higgs couplings: Stau annihilation into $\\hhiggs\\hhiggs$, $\\hhiggs\\Hhiggs$, and $\\Hhiggs\\Hhiggs$ can become very efficient (if kinematically allowed) so that $\\Ystau$ can decrease to values well below $10^{-15}$. The scalar nature of $\\stauone$ allows those parameters to enter directly into the annihilation cross sections. This mechanism has no analogue in calculations of the thermal relic density of the lightest neutralino $\\neutralino$. The stau--Higgs couplings are crucial also for the second $\\Ystau$ reduction mechanism identified in this work: Even for moderate values of $\\tanb$, we find that staus can annihilate very efficiently into a $\\bquark\\antibquark$ pair via $s$-channel exchange of the heavy CP-even Higgs boson $\\Hhiggs$ provided the MSSM spectrum exhibits the resonance condition $2\\mstauone\\simeq\\mH$. We have shown explicitly that the associated $\\Ystau$ values can be below $10^{-15}$ as well. This mechanism is similar to the one that leads to the reduction of the $\\neutralino$ density in the Higgs funnel region in which neutralino annihilation proceeds at the resonance of the CP-odd Higgs boson $\\Ahiggs$. We have worked with an effective low energy version of the MSSM to investigate the $\\thetastau$-dependence of $\\Ystau$ and the two $\\Ystau$-reduction mechanisms in a controlled way. In addition, we have shown that the considered effects can be accommodated also with restrictive assumptions on the soft-SUSY breaking sector at a high scale. Within the CMSSM, we encounter both mechanisms each of which leading to $\\Ystau\\simeq 2\\times 10^{-15}$ in two distinct regions of a single $(\\monetwo,\\,\\mzero)$ plane. We have discussed possibilities to probe the viability of the presented $\\Ystau$-reduction mechanisms at colliders. While a $\\mH$ measurement pointing to $\\mH\\simeq 2\\mstauone$ would support resonant primordial stau annihilation, studies of Higgs boson production in association with staus, $\\positron\\electron\\,(\\gamma\\gamma) \\to\\stauone\\stauone^*\\hhiggs,\\stauone\\stauone^*\\Hhiggs$ could allow for an experimental determination of the relevant stau--Higgs couplings, for example, at the ILC. Moreover, we have outlined that associated $\\bquark\\antibquark\\hhiggs/\\Hhiggs$ production with $\\hhiggs/\\Hhiggs\\to\\stauone\\stauone^*$ has the potential to allow for a determination of both $\\mh$ and $\\mH$ at the LHC if a SUSY scenario with large $\\tanb$ and large stau--Higgs couplings is realized. With the obtained small $\\Ystau$ values, even the restrictive constraints associated with CBBN could be respected so that attractive gravitino dark matter scenarios could be revived to be cosmologically viable even for a standard cosmological history. Within this class of models, collider evidence for supergravity, for the gravitino being the LSP, and for high values of the reheating temperatures of up to $10^9\\,\\GeV$ is conceivable, which could thereby accommodate simultaneously the explanation of the cosmic baryon asymmetry provided by thermal leptogenesis and the hypothesis of thermally produced gravitinos being the dark matter in our Universe. \\bigskip {\\bf Acknowledgments} -- We are grateful to T.~Hahn, J.-L.~Kneur, and A.~Pukhov for patient correspondence on their computer codes. Furthermore, we are grateful to T.~Plehn, M.~Pospelov, and Y.Y.Y.~Wong for valuable discussions. This research was supported in part by the DFG cluster of excellence ``Origin and Structure of the Universe.'' \\bigskip {\\bf Note added} -- Ref.~\\cite{Ratz:2008qh}, in which the potential suppression in the stau yield $\\Ystau$ due to an enhanced annihilation into $\\hhiggs\\hhiggs$ final states is also studied, appeared as this work was being finalized. This paper provides analytic approximations for the stau annihilation cross section into $\\hhiggs\\hhiggs$ and for the associated yield. In addition, results of numerical studies within the CMSSM, the NUHM, and a scenario with non-universal gaugino masses are presented that exhibit parameter regions with extremely small $\\Ystau$ values. In our work also enhanced stau annihilation into $\\hhiggs\\Hhiggs$ and into $\\Hhiggs\\Hhiggs$ and stau annihilation at the $\\Hhiggs$ resonance, which were not considered in~\\cite{Ratz:2008qh}, are discussed. In addition, our work provides a systematic investigation based on a complete set of stau annihilation channels, an outline of the way in which the mechanisms leading to the suppression of $\\Ystau$ can be probed at collider experiments, and a thorough presentation of the potential implications for gravitino dark matter scenarios." }, "0808/0808.0184_arXiv.txt": { "abstract": "TeV-blazars are known as prominent non-thermal emitters across the entire electromagnetic spectrum with their photon power peaking in the X-ray and TeV-band. If distant, absorption of $\\gamma$-ray photons by the extragalactic background light (EBL) alters the intrinsic TeV spectral shape, thereby affecting the overall interpretation. Suzaku observations for two of the more distant TeV-blazars known to date, 1ES~1101-232 and 1ES~1553+113, were carried out in May and July 2006, respectively, including a quasi-simultaneous coverage with the state of the art Cherenkov telescope facilities. We report on the resulting data sets with emphasis on the X-ray band, and set into context to their historical behaviour. During our campaign, we did not detect any significant X-ray or $\\gamma$--ray variability. 1ES~1101-232 was found in a quiescent state with the lowest X-ray flux ever measured. The combined XIS and HXD PIN data for 1ES~1101-232 and 1ES~1553+113 clearly indicate spectral curvature up to the highest hard X-ray data point ($\\sim 30$ keV), manifesting as softening with increasing energy. We describe this spectral shape by either a broken power law or a log-parabolic fit with equal statistical goodness of fits. The combined 1ES 1553+113 very high energy spectrum ($90-500$~GeV) did not show any significant changes with respect to earlier observations. The resulting contemporaneous broadband spectral energy distributions of both TeV-blazars are discussed in view of implications for intrinsic blazar parameter values, taking into account the $\\gamma$-ray absorption in the EBL. ", "introduction": "Blazars are among the most extreme sources in the high energy sky. They constitute a subclass of the jet-dominated radio-loud population of active galactic nuclei (AGN), which in turn are sub-divided into flat spectrum radio quasars (FSRQ) and BL Lac objects where the absence or depression of strong emission lines characterizes the latter. Their observational properties include non-thermal continuum emission and irregular variability across the whole electromagnetic band, ranging from subhours at gamma-ray energies to week-long time scales in the radio band, often high optical and radio polarization, and a core-dominated radio morphology. In some sources, superluminal motion has been detected, indicating a relativistically enhanced emission region close to the line-of-sight \\cite{Blandford78}. As a consequence of this, any jet emission is highly beamed. The spectral energy distribution (SED) of blazars shows two broad peaks in the $\\nu F_\\nu$ representation. The low energy hump is generally agreed to stem from synchrotron radiation of relativistic electrons and/or positrons in the jet whereas the origin of the high energy one is still under debate. Leptonic models explain the complete non-thermal SED as synchrotron and inverse Compton emission from relativistic electrons or pairs which upscatter their self-produced synchrotron radiation or external photon fields. Hadronic models produce the high energy peak either via interactions of relativistic protons with matter \\cite{Pohl00}, ambient photons \\cite[e.g.][]{Mannheim93}, magnetic fields \\cite{Aharonian00}, and/or both, magnetic fields and photons \\cite[e.g.][]{Muecke01,Muecke03}. Almost all of the 19 blazars (except for BL Lacertae \\cite{Albert07c} and 3C~279 \\cite{Teshima07}) detected reliably to date at the TeV energy belong to the subclass of high-frequency peaked BL Lacs (HBLs), with their two $\\nu F_\\nu$ peaks typically at UV/X-rays and TeV-energies. Until recently blazars have been detected preferentially in their flare state, owing to instrumental limitations. With the contemporary Cherenkov telescopes (such as H.E.S.S., MAGIC, VERITAS) it is possible to study also low activity states (subsequently referred to \"quiet state\"). TeV-blazars have also been used successfully as a powerful tool to probe the extragalactic background light (EBL) at IR/optical energies via the absorption of $\\gamma$-rays in the cosmic diffuse radiation field. The most stringent constraints on the EBL density at $\\sim 2\\mu$m stem from observations with the H.E.S.S. telescope system of the high-redshift, hard-spectrum TeV-blazar 1ES~1101-232 \\cite{Aharonian06a}. The most distant (redshift $z>0.1$) TeV-blazars established to date are 1ES~1101-232 ($z=0.186$), 1ES~1218+304 ($z=0.182$), 1ES~1011+496 ($z=0.212$), 1ES~0347-121 ($z=0.185$), 1ES~0229+200 ($z=0.140$), H~2356-309 ($z=0.165$), H~1426+428 ($z=0.129$), PKS~2155-304 ($z=0.116$), 1ES 1553+113 \\cite[$z>0.25$:][]{Perlman96,Falomo90,Aharonian06b}. Extreme HBLs, as is the case for most very high energy (VHE) blazars, are characterized by their spectral extension to extremely high particle energies. TeV-blazars have been regularly observed in the past by X-ray instruments, however, not many have been detected so far in the hard X-ray range where the highest energy particles leave their imprints. The sensitivity reached by the hard X-ray detector HXD onboard the Suzaku satellite allowed for the first time to study in detail possible new spectral features and curvature at $>10$~keV, even in a blazar quiet state, which may help understanding the properties and origin of the VHE emission. We report here on our $\\sim 50$~ksec Suzaku observations on each of two more distant TeV-blazars known to date: 1ES~1101-232 and 1ES~1553+113. These observations were covered by quasi-simultaneous observations with the H.E.S.S. telescope system and the MAGIC telescope. None of these sources have shown in the past significant variability in the TeV-band \\cite{Aharonian06a,Aharonian06b}, likely owing to their low flux close to the detection limit of the TeV facilities. The paper is organized as follows: We provide a brief overview of both TeV-blazars in Sect.~2. Sect.~3 is devoted to a technical description of our Suzaku data analysis and its results. In Sect.~4 we report on the quasi-simultanuous multiwavelength coverage during the Suzaku observations of 1ES 1101-232 and 1ES 1553+113. In Sect.~5 we discuss our results in particular in context to their historical behaviour. Sect.~6 summarizes the main results of this work. ", "conclusions": "\\subsection{Historical X-ray behaviour of 1ES 1101-232 and 1ES 1553+113} 1ES 1101-232 has been detected during the past 15 years by a range of X-ray instruments, both covering the soft (ROSAT, XMM) and extending into the hard X-ray range (BeppoSAX, RXTE, SWIFT) at various sensitivity levels. Except for the continuous monitoring by RXTE's ASM and the 19.6~ksec monitoring by XMM \\cite{Aharonian07a}, most observations were carried out in snapshot mode, with each observation encompassing typically 1-5 ksec. The continuous Suzaku observation presented here constitutes therefore the first long ($>30$~ksec) high-sensitivity X-ray data set of this source. None of the past observations showed significant short time variability ($<$hour). Although the continuous mode observations of Suzaku generally allow to probe intraday variability, we find no significant variability of this time scale during our observations. Year-by-year variations as derived from the historical X-ray light curve are less than a factor $\\sim 2$. The highest ever measured X-ray flux level has been reported for 2005 from the RXTE-PCA measurements \\cite{Aharonian07a}. There, in eleven consecutive nights the source was observed each night for $\\sim 10$~ksec. No hints for strong flux variability were found, and the AGN was categorized to be in a non-flaring state. If this is the case, 1ES 1101-232 was likely never observed in a flaring state during the past 15 years. Only 4 months later SWIFT saw the source in a smooth decline of flux level with no sub-hour variations \\cite[e.g.][]{Massa08}. Our Suzaku observations, made approximately $10$ months later, found the lowest ever measured X-ray flux from 1ES 1101-232. So far, essentially all sensitive X-ray observations of 1ES 1101-232 prefer a curved or broken power-law representation instead a simple power-law \\cite[e.g.,][]{Perlman05}, independent of flux level and - as demonstrated in this work - extending into the hard X-ray range. If interpreted as signatures of the particle energy gain process, one may conclude that acceleration takes place at all flux levels observed so far. In all cases there were no indications for a significant amount of X-ray absorbing material in the line-of-sight beyond the Galactic absorption column density. The average spectrum during our Suzaku observations was quite similar to the one from the 20\\% higher flux XMM observations in 2001 \\cite[$\\Gamma\\sim2.1-2.4$:][]{Perlman05}, where also no significant variability was noticed. The 1997/98 BeppoSAX observations \\cite{Wolter00} had a similar value of the break energy at a factor $\\sim 1.3-1.9$ higher flux level. The flux reported from the 2004 observations with XMM was comparable to the one measured by BeppoSAX in 1997, however, the spectral break shifted to lower energies by a factor $\\sim 1.2$. Because of the lacking soft X-ray sensitivity, the PCA coverage of 1ES 1101-232 in 2005 revealed a break energy at much higher energies ($\\sim 8$~keV). If the 1-8~keV X-ray spectral index is compared to the observed flux level at each observation, no statistically convincing indication for a flux-index correlation is found (Spearman rank correlation coefficient $R_s=-0.34$ with chance probability $P_c=0.33$, Kendall's $\\tau_K=-0.30$ with $P_c=0.23$). Comparing, e.g., the XMM measurements carried out in 2001 and 2004, no significant spectral changes within the uncertainties occured despite a factor $\\sim 1.6$ higher flux in 2004. On the other side, the BeppoSAX observations revealed a spectral softening from the higher flux state in 1997 to the lower flux state in 1998. When compared to the high flux RXTE data from 2005 a noticable softening occured despite the overall high flux level. If following the evolution of the spectrum versus flux, no clear spectral hysteresis could be found. In summary, it seems that spectral changes observed from annual visits of this source do not follow particular patterns or relations with respect to flux changes. The overall spectral changes lie within a rather narrow range between $\\sim 2$ and 2.5 for the 1-8~keV photon spectral index. The measured so far comparably long X-ray variability time scale in 1ES 1101-232 may imply that this source was observed only in the quiescence over the last $\\sim 15$~years. If one assumes the measured activity states from one year to the other to be not causally related, the annual measurements may then be treated as an independent ensemble of measurements from an HBL. In particular, it seems reasonable to probe the behaviour in the peak luminosity-peak energy diagram within the available HBL range. For a robust determination of the peak energy, low energy data, preferably in the IR/optical/UV range, are necessary. Giommi et al. (2005) used optical data from the literature in combination with the BeppoSAX measurements, while both XMM pointings had the advantage of simultaneous optical coverage with the Optical Monitor (OM) onboard the satellite, and SWIFT the onboard UVOT instrument. We did not find any convincing anti-correlation between peak energy and peak flux within the available 6 ensemble members ($R_s=-0.6$ with $P_c=0.21$, $\\tau_K=-0.33$ with $P_c=0.35$). 1ES~1553+113 is known as a generally X-ray soft, bright HBL with only modest X-ray variability (factor $\\leq 3.5$). SWIFT in 2005 \\cite{Trama07,Massa08} achieved a detection of this source up to only 8~keV, despite its high flux level in October 2005 that was reached within half a year from a 3.5 times lower flux level. BeppoSAX in 1998 \\cite{Donato05} saw it only with the LECS/MECS at a relatively low flux level, no detection with the PDS was possible. With our Suzaku observation, 1ES~1553+113's spectrum could finally be measured up to 30~keV while descending from the 2005 high flux level. Like for 1ES 1101-232, past X-ray observations encompass mostly snapshot observations of typically 3-10~ksec each. RXTE's observation campaign in 2003 \\cite{Osterman06} involved visits to the source about 3 times per day for $\\sim 3$~ksec, for a total period of 21 days. During this time a very smooth rise in flux up to 3 times the lowest flux level was observed. No hints of subhour variations were recorded in any of the past observations. With our $\\sim 50$~ksec continuous Suzaku measurements we tested for intraday variability, but obtained a null result also at this intermediate flux level which was quite comparable to the 2001 flux recorded by XMM \\cite{Perlman05}. No significant differences were found between the spectra reported from the 2001 and 2006 observations: the average spectrum during our Suzaku observations was, within the uncertainties, in good agreement to the one from the XMM observations in 2001 \\cite[$\\Gamma\\sim2.2-2.4$: ][]{Perlman05}. The surprising spectral stability of this source holds even for the low flux state recorded with RXTE in 2003 \\cite{Osterman06}, despite the factor $\\sim$~5 difference in flux. Even within the tripling of the flux during the PCA observations spectral changes were not significantly beyond their uncertainties. In particular, any systematic spectral trend with changing flux is not apparent, neither on a year-by-year time scale nor within the rising part of the RXTE monitoring data within their admittedly large uncertainties: as is the case for 1ES 1101-232, the photon index above $3$~keV typically lingered around 2 and 2.5 even during flux increases by a factor $\\sim 2$ or more. Those modest spectral changes together with the smoothness of the long-term light curve may suggest that explosive events are among the least likely scenarios to account for the observed flux variations here. The 1998 BeppoSAX observations \\cite{Donato05} had a lower break energy ($\\sim1$~keV) at a factor $\\sim 2.5$ lower flux level as compared to the 2006 Suzaku observations. SWIFT detected 1ES 1553+113 in both, a rather low flux state (April 2005) and relatively high flux level (October 2005), but with a surprisingly stable peak energy of $\\sim 0.4-0.5$~keV in the XRT spectrum \\cite[e.g.][]{Massa08}, $\\sim 3$ times lower than the Suzaku data from 2006 imply. The flux reached during the April 2005 observations with SWIFT was comparable to the one measured by BeppoSAX in 1998, however the peak energy from the XRT data shifted to lower energies by a factor $\\sim 2.5$. No break energies were reported from the 2003 RXTE/PCA data \\cite{Osterman06}. Following the argumentation above we probe here the dependence between peak luminosity and peak energy also for 1ES 1553+113, and utilize preferably data from the IR/optical/UV range to complement the X-ray band measurements. Among the past observations only the XMM and SWIFT pointings were quasi-simultaneous measurements with optical/UV available, namely OM, UVOT and ROTSE data, respectively. We use here our Suzaku observations with the data from the quasi-simultaneous optical coverage by the KVA optical telescope for a combined X-ray - optical spectral fit. A log-parabolic shape $dN/dE \\propto E^{[-\\Gamma-\\beta\\log(E)]}$ was used to determine the peak energy to $\\sim 0.01$~keV with parameters $\\beta=0.075\\pm 0.001$ and $\\Gamma=2.298\\pm 0.003$. Note, however, that this representation does only give a $\\chi_{\\rm red}^2 \\approx 5.7$ due to the large weight given to the only optical point (that has a small uncertainty) with respect to the many X-ray data points. More lower energy points would clarify this. We did not use the peak values given for the 2003 RXTE campaign for this undertaking to assure for a sample with equally determined peak energies: the cm radio data from Osterman et al. (2006) for determining the peak energy would give too much weight towards the longer wavelengths when compared to a procedure where only optical/UV in combination with the X-ray data are used. The peak energies from the UVOT measurements range from $\\sim 0.05$ to 0.02~keV from the high to the low flux state, respectively. The latter is in very good agreement with the values derived from the XMM 2001 observations. Also for 1ES 1553+113 any convincing systematic correlation between peak energy and peak flux within the very limited sample of 4 ensemble members is lacking evidence ($R_s=0.39$ with $P_c=0.61$, $\\tau_K=0.55$ with $P_c=0.26$). This may indicate that particle acceleration up to the maximum energy is not preferentially limited by Compton losses if beaming does not change significantly. Indeed, because of the lack of sufficiently dense radiation fields in HBL-type AGN, synchrotron losses may dominate the radiative particle loss channel. The absence of a link between the peak luminosity and peak energy can then be understood if either an increase of the jet's synchrotron radiation field is not connected to an increase of the magnetic field strength alone on year time scales, or expansion losses and/or escape from the emission region sets a boundary to the maximum electron energy at the so far observed activity states in 1ES~1101-232 and 1ES~1553+113. \\subsection{Spectral energy distributions} The overall SED of 1ES~1101-232, including the Suzaku spectrum obtained by us and the simultaneous H.E.S.S. data is shown in Fig.~\\ref{fig3}. If corrected by the minimal EBL \\cite[P0.45 from][]{Aharonian06a}, comparable power is emitted at $\\gamma$-rays than at the synchrotron hump with a $\\gamma$-ray peak beyond TeVs at the time of the quasi-simultaneous Suzaku -- TeV observations. In the framework of a simple one-zone SSC model a (assumed uniform) field strength of sub-equipartition strength $B^2 \\propto (\\nu F_{\\nu,IC})/(\\nu F_{\\nu,syn})$ can then be deduced. Alternatively, further target photon fields in the vicinity of the $\\gamma$-ray emission zone can lead to an enhanced $\\gamma$-ray photon output. In hadronic blazar emission models TeV-emission is the result of either reprocessed cascade emission initiated by pairs and/or photons produced in photomeson interactions, or by proton and/or $\\pi^{\\pm}/\\mu^{\\pm}$-synchrotron radiation \\cite[e.g.][]{Muecke01,Muecke03}. In environments of weak ambient photon fields proton-photon interactions constitute a correspondingly sub-dominant contribution to the $\\gamma$-ray component, which then leaves proton synchrotron (or curvature) radiation as the main $\\gamma$-ray producer. \\clearpage \\begin{figure}[t] \\resizebox{\\hsize}{!}{\\includegraphics{f9.eps}} \\caption{Broadband SED of 1ES~1101-232. The observed H.E.S.S. data were de-absorbed using the minimal EBL \\cite[P0.45 from][]{Aharonian06a}. Present Suzaku data and the simultaneous de-absorbed July 2006 H.E.S.S. data are indicated in red. Blue points (XMM, RXTE, HESS) indicate data from 2004/05 \\cite{Aharonian07a}, while historical data collected from the literature are in black open symbols, and light grey \\cite[BeppoSAX:][]{Wolter00}. The SWIFT data \\cite[][not shown here]{Massa08} are compatible with the XMM flux level.} \\label{fig3} \\end{figure} \\clearpage TeV observations of 1ES~1101-232 were used previously \\cite{Aharonian06a} to set the most stringent limits on the EBL at $\\sim 0.7-4\\mu$m using spectral considerations \\cite{Aharonian06a}. These limits could not be improved owing to the limited quality from the low flux TeV-data from 2006. The overall SED of 1ES~1553+113 with the contemporaneous Suzaku, VHE (H.E.S.S., MAGIC) and optical (KVA) data is shown in Fig.~\\ref{fig10}, complemented with historical ones from the literature. Using a minimal EBL (P0.45) for de-absorption, approximately equal or more power is emitted at $\\gamma$-rays than at the synchrotron hump with a $\\gamma$-ray peak at sub-TeVs. In the framework of a simple one-zone SSC model, this ratio indicates sub-equipartition field strengths in the emission region. Furthermore, if the same electrons emit TeV- and X-ray photons co-spatially, then the intrinsic spectral index at VHEs must not be harder than at X-ray energies. For an EBL density at the galaxy counts flux level this is the case for $z \\geq 0.4$. Alternatively, external photon fields may contribute to the target photon fields for inverse Compton scattering, or hadronic emission models may be at work here. \\clearpage \\begin{figure}[t] \\resizebox{\\hsize}{!}{\\includegraphics{f10.eps}} \\caption{Broadband non-simultaneous SED of 1ES~1553+113. The quasi-simultaneous data points (KVA, Suzaku, H.E.S.S., MAGIC) data are indicated as (red) squares. The green/grey line respresents the highest flux level observed by SWIFT \\cite{Massa08} of this source. The $\\gamma$-ray data have been de-absorbed for a source redshift of $z=0.3$ using the minimal EBL \\cite[P0.45 from][]{Aharonian06a}.} \\label{fig10} \\end{figure} \\clearpage" }, "0808/0808.3843_arXiv.txt": { "abstract": "Radio observations from decimetric to submillimetric wavelengths are now a basic tool for the investigation of comets. Spectroscopic observations allow us i) to monitor the gas production rate of the comets, by directly observing the water molecule, or by observing secondary products (e.g., the OH radical) or minor species (e.g., HCN); ii) to investigate the chemical composition of comets; iii) to probe the physical conditions of cometary atmospheres: kinetic temperature and expansion velocity. Continuum observations probe large-size dust particles and (for the largest objects) cometary nuclei. Comets are classified from their orbital characteristics into two separate classes: i) nearly-isotropic, mainly long-period comets and ii) ecliptic, short-period comets, the so-called Jupiter-family comets. These two classes apparently come from two different reservoirs, respectively the Oort cloud and the trans-Neptunian scattered disc. Due to their different history and --- possibly --- their different origin, they may have different chemical and physical properties that are worth being investigated. The present article reviews the contribution of radio observations to our knowledge of the Jupiter-family comets (JFCs). The difficulty of such a study is the commonly low gas and dust productions of these comets. Long-period, nearly-isotropic comets from the Oort cloud are better known from Earth-based observations. On the other hand, Jupiter-family comets are more easily accessed by space missions. However, unique opportunities to observe Jupiter-family comets are offered when these objects come by chance close to the Earth (like 73P/Schwassmann-Wachmann 3 in 2006), or when they exhibit unexpected outbursts (as did 17P/Holmes in 2007). About a dozen JFCs were successfully observed by radio techniques up to now. Four to ten molecules were detected in five of them. No obvious evidence for different properties between JFCs and other families of comets is found, as far as radio observations are concerned. ", "introduction": "\\label{sect-intro} Jupiter-family comets (JFCs, aka \\textit{ecliptic comets}) are short-period, low-inclination comets likely to undergo orbital perturbations by Jupiter. We adopt here the definition based on the Tisserand invariant with respect to Jupiter $T_J$: JFCs are comets for which $2 < T_J < 3$ \\citep{levi:1996}. Special cases are 2P/Encke which slightly exceeds the $T_J = 3$ limit and for this reason sometimes classified as a \\textit{Encke-type} comet, and 29P/Schwassmann-Wachmann 1 with $T_J = 2.99$, alternatively classified as a \\textit{Centaur}. JFCs are believed to come from the trans-Neptunian scattered disc. They contrast with the \\textit{nearly isotropic comets}, which comprise long-period comets as well as short-period comets (the so-called \\textit{Halley type comets}), supposed to come from the Oort cloud. Comets certainly did not form in these trans-Neptunian reservoirs. Their real sites of formation and their orbital evolution are still highly debated topics. (For reviews on cometary families and their dynamical evolution, see \\citet{levi:1996, fern:2008, lowr+:2008, morb:2008, morb+:2008, mars:2008}). Having different dynamical histories and --- possibly --- different origins, these distinct classes of comets may have different chemical and physical properties that are worth being investigated. \\begin{table*} \\caption{Remote sensing observing conditions for a selection of comets} \\label{table-remote} \\begin{center} \\begin{tabular*}{\\hsize}[]{@{\\extracolsep{\\fill}}llcccc} \\hline comet & date & $r_h$ & $\\Delta$ & $Q_\\mathrm{H_2O}$ & $FM$ \\\\ & & [AU] & [AU] & [$10^{28}$ s$^{-1}$] \\\\ \\hline \\multicolumn{5}{l}{\\textit{Nearly-isotropic comets}} \\\\ C/1986 P1 (Wilson) & May 1987 & 1.3 & 1.0 & 12 & 12 \\\\ C/1989 X1 (Austin) & April 1990 & 1.2 & 0.25 & 2.5 & 10 \\\\ C/1990 K1 (Levy) & August 1990 & 1.3 & 0.45 & 25 & 55 \\\\ C/1996 B2 (Hyakutake) & March 1996 & 1.1 & 0.11 & 25 & 225 \\\\ C/1995 O1 (Hale-Bopp) & April 1997 & 0.9 & 1.4 &1000 & 700 \\\\ C/1999 S4 (LINEAR) & July 2000 & 0.77& 0.38 & 10 & 25 \\\\ C/1999 T1 (McNaught-Hartley) & December 2000 & 1.2 & 1.6 & 10 & 6 \\\\ C/2001 A2 (LINEAR) & June 2001 & 1.0 & 0.24 & 10 & 40 \\\\ C/2000 WM$_1$ (LINEAR) & December 2001 & 1.2 & 0.32 & 4 & 12 \\\\ C/2001 Q4 (NEAT) & May 2004 & 1.0 & 0.32 & 20 & 60 \\\\ C/2002 T7 (LINEAR) & May 2004 & 0.8 & 0.27 & 10 & 25 \\\\ C/2003 K4 (LINEAR) & December 2004 & 1.4 & 1.2 & 15 & 12 \\\\ C/2004 Q2 (Machholz) & January 2005 & 1.2 & 0.35 & 25 & 70 \\\\ \\hline \\multicolumn{5}{l}{\\textit{Halley-type comets}} \\\\ 1P/Halley & January 1986 & 0.7 & 1.5 & 120 & 80 \\\\ 109P/Swift-Tuttle & November 1992 & 1.0 & 1.3 & 50 & 40 \\\\ 153P/2002 C1 (Ikeya-Zhang) & April 2002 & 1.0 & 0.40 & 25 & 60 \\\\ 8P/Tuttle & January 2008 & 1.1 & 0.25 & 3 & 12 \\\\ \\hline \\multicolumn{5}{l}{\\textit{Jupiter-family comets}} \\\\ 22P/Kopff & April 1996 & 1.7 & 0.9 & 3.5 & 4 \\\\ 21P/Giacobini-Zinner & October 1998 & 1.2 & 1.0 & 3 & 3 \\\\ 19P/Borrelly & September 2001 & 1.36 & 1.47 & 3 & 2 \\\\ 2P/Encke & November 2003 & 1.0 & 0.25 & 0.5 & 2 \\\\ 9P/Tempel~1 & July 2005 & 1.5 & 0.77 & 1 & 1.3 \\\\ 73P/Schwassmann-Wachmann~3 & May 2006 & 1.0 & 0.08 & 2 & 25 \\\\ 17P/Holmes & October 2007 & 2.4 & 1.6 & $>100$ & $>60$ \\\\ 67P/Churyumov-Gerasimenko & Mars 2009 & 1.24 & 1.7 & 1 & 0.6 \\\\ 103P/Hartley~2 & October 2010 & 1.1 & 0.12 & 1.2 & 10 \\\\ 45P/Honda-Mrkos-Pajdu\\v{s}\\'akov\\'a & August 2011 & 1.0 & 0.06 & 0.5 & 8 \\\\ \\hline \\end{tabular*} \\end{center} Date is for best observing conditions. \\\\ $r_h$ = distance to Sun and $\\Delta$ = distance to Earth at that date. \\\\ $Q_\\mathrm{H_2O}$ = water production rate at that date. \\\\ $FM = Q_\\mathrm{H_2O}/\\Delta$ = \\textit{figure of merit}, roughly proportional to the signal of cometary molecules. \\end{table*} JFCs and other comets are far from being equally well observed. To show this for Earth-based observations, we will use the \\textit{figure of merit} parameter $FM = Q_\\mathrm{H_2O}/\\Delta$, where $Q_\\mathrm{H_2O}$ is the water production rate in units of $10^{28}$ s$^{-1}$ and $\\Delta$ is the distance to the observer in AU. (Note that this parameter differs from the \\textit{figure of merit} introduced by \\citet{mumm+:2002}, which includes a dependency on the distance to the Sun.) This parameter is roughly proportional to the expected signal intensity (for comets at heliocentric distances of the order of 1~AU), and allows us to evaluate and compare the observability of comets. Table~\\ref{table-remote} gives the figures of merit for Earth-based observations of recent long-period comets, as well as for recent and future returns of short-period comets. One can see that unexpected, long-period comets from the nearly isotropic class of comets and Halley-type comets offer much better opportunities than short-period comets. Indeed, the two best comets in the last twenty years were C/1996 B2 (Hyakutake) and C/1995 O1 (Hale-Bopp); unprecedented spectroscopic observations leaded to the identification of many molecules for the first time in these two objects \\citep{bock+:2005}. All JFCs are comets with low water production rates $Q_\\mathrm{H_2O}$ of at most a few 10$^{28}$~s$^{-1}$ at perihelion. The best observing conditions occur for these comets which make a close approach to the Earth (i.e., for $\\Delta$ significantly smaller than 1~AU). This was recently the case for 73P/Schwassmann-Wachmann 3 (minimum geocentric distance $\\Delta = 0.08$ AU in May 2006) and it will also happen for 103P/Hartley~2 in the near future ($\\Delta = 0.12$ AU in October 2010). But the \\textit{figure of merit} of such JFCs is still lower than that of some long-period comets for which $FM$ could exceed 50. Other unique opportunities also occur when JFCs exhibit unexpected outbursts during which the gas production rate may be increased by several orders of magnitude. This was recently the case for 17P/Holmes, but such events are rare. On the other hand, short-period comets, with their predicted returns, are the only practicable targets for space missions, which are not yet versatile enough to accommodate unexpected comets. Flybys at a low velocity and rendezvous are only possible for ecliptic comets, due to the energy limitations of current space technology. 1P/Halley is the only explored comet which does not belong to the Jupiter family: the price to pay was a very high flyby velocity ($\\approx 70$ km~s$^{-1}$). Spectroscopy at radio wavelengths is now a basic tool for the investigation of comets. It allows us: \\begin{itemize} \\item to monitor the gas production rate of the comets, by directly observing the water molecule, or by observing secondary products (e.g., the OH radical) or minor species (e.g., HCN); \\item to investigate the chemical composition of comets; \\item to probe the temperature of cometary atmospheres by observing simultaneously several rotational lines of the same molecule; \\item to investigate the expansion velocity and kinematics of cometary atmospheres from the observation of line shapes. \\end{itemize} Continuum observations of comets in the radio spectral range are also well suited for studying the nucleus and the dust coma via their thermal radiation. Since millimetric and submillimetric radiation can only be efficiently radiated from large particles ($\\gtrsim 1$ mm) which comprise most of the dust mass, this technique is a useful probe of the total dust mass production and of the size distribution of the large grains. The present article reviews the contribution of radio observations to our knowledge of JFCs. The outcome of radar experiments \\citep{harm+:2005}, which probe the nucleus and large dust particles, is beyond the scope of this paper. ", "conclusions": "Radio observations of JFCs (see Table~\\ref{table-summary} for a summary) are still sparse. Up to now, successful radio observations of JFCs have been limited to a dozen objects, mainly restricted to exceptional comets (comets which made a close approach to Earth and/or which showed an outburst), or to the targets of space missions for which a special effort was made. New radio instruments are to be soon available for the observations of comets with an increased sensitivity: \\begin{itemize} \\item the Large Millimeter Telescope (LMT), under completion in Mexico \\citep{irvi-schl:2005}; with its 50-m antenna, it will be soon the largest millimetric telescope of its category; \\item the Herschel Space Observatory, to be launched in 2009 \\citep{crov:2005}; \\item the Atacama Large Millimetre Array (ALMA) in Chili \\citep{bive:2005, bock:2008}; \\item the Square Kilometre Array (SKA), which will operate at decimetric-centimetric wavelengths \\citep{butl+:2004}; \\item the MIRO instrument on Rosetta, a 30-cm diameter radio telescope to observe dedicated molecular radio lines in situ in comet 67P/Churyumov-Gerasimenko, as well as the continuum emission of its nucleus \\citep{gulk+:2007}. \\end{itemize} Specific opportunities to observe JFCs with a high $FM$ will occur in the near future (Table~\\ref{table-remote}): \\begin{itemize} \\item 103P/Hartley~2 in October 2010 (with a flyby of the redirected Deep Impact spacecraft, renamed as the EPOXI mission); \\item 45P/Honda-Mrkos-Pajdu\\v{s}\\'akov\\'a in August 2011. \\end{itemize} It will thus be possible to significantly increase the sample of JFCs with known molecular composition, and to pursue the comparative study of the different cometary families. Isotopic measurements in cometary volatiles are also important diagnostics on the origin of cometary material. Most measurements have been made using millimetre or submillimetre spectroscopy. However, if one excepts the Jupiter-family comet 17P/Holmes (Section~\\ref{17P}), the capabilities of ground and space-based instrumentation have limited the investigations to a few bright long-period comets, as reviewed in \\citet{altw-bock:2003} and \\citet{bock+:2005} and reported in \\citet{bive+:2007-PASS}. Isotopic ratios could be different in JFCs and Oort cloud comets if the two populations formed at different places or different times in the solar nebula. Though the LMT promises to be very useful for the study of the molecular composition of JFCs, isotopic investigations with this telescope will be limited to exceptionally bright comets \\citep{irvi-schl:2002}. The HIFI instrument of the Herschel Space Observatory is to provide the opportunity to detect the 1$_{10}$--1$_{01}$ line of HDO at 509.3~GHz in comet 103P/Hartley 2, and to observe simultaneously several H$_2$O lines for an accurate determination of the D/H ratio in water. ALMA will allow the detection of HDO at 465 and 894 GHz in short-period comets with $FM > 5$, and DCN in comets with $FM > 10$. \\balance \\footnotesize{" }, "0808/0808.1594_arXiv.txt": { "abstract": "Using the recently upgraded Long Baseline Array, we have measured the trigonometric parallax of PSR J0437--4715 to better than 1\\% precision, the most precise pulsar distance determination made to date. Comparing this VLBI distance measurement to the kinematic distance obtained from pulsar timing, which is calculated from the pulsar's proper motion and apparent rate of change of orbital period, gives a precise limit on the unmodeled relative acceleration between the Solar System and PSR J0437--4715, which can be used in a variety of applications. Firstly, it shows that Newton's gravitational constant $G$ is stable with time ($\\Gdot/G = (-5 \\pm 26) \\times 10^{-13}$\\ yr$^{-1}$, 95\\% confidence). Secondly, if a stochastic gravitational wave background existed at the currently quoted limit, this null result would fail $\\sim\\!50$\\% of the time. Thirdly, it excludes Jupiter-mass planets within 226 AU of the Sun in 50\\% of the sky (95\\% confidence). Finally, the $\\sim\\!1$\\% agreement of the parallax and orbital period derivative distances provides a fundamental confirmation of the parallax distance method upon which all astronomical distances are based. ", "introduction": "PSR J0437--4715 has been observed intensively since its discovery by \\citet{johnston93a}. It is the brightest and nearest observed millisecond pulsar, and has also been studied in the optical \\citep{bell93a}, ultraviolet \\citep{kargaltsev04a}, and X--ray \\citep{zavlin02a} bands. The high rotational stability and close proximity of this pulsar--white dwarf binary system make it an excellent probe of General Relativity (GR) and alternative forms of gravitational theories. The measurement of its Shapiro delay by \\citet{van-straten01a}, where the radio waves from the pulsar are delayed as they pass through the gravitational potential of the companion, is one such test which has shown consistency with GR predictions. The search for the low frequency stochastic gravitational wave background (GWB) using pulsar timing arrays \\citep[e.g.][]{jenet05a} is another test of GR which is facilitated in part by timing of J0437--4715. Although variation of Newton's gravitational constant $G$\\ with time is forbidden in GR, this is not required in alternate formalisms of gravity \\cite[e.g.][]{brans61a}. \\citet{verbiest08a} have measured the rate of change of orbital period \\pbdot\\ in the J0437--4715 system to better than 2\\% precision and shown that the agreement of the derived kinematic distance with the parallax distance derived from timing limits the time variation of $G$\\ to less than 3 parts in $10^{11}$. In this Letter, we present a new Very Long Baseline Interferometry (VLBI) determination of the position, proper motion, and parallax of PSR J0437--4715 and show that this improves the previously published \\Gdot\\ limit derived from this system by an order of magnitude, approaching the best published limit ($\\Gdot/G = (4\\pm9)\\times 10^{-13}$\\ yr$^{-1}$), which is derived from Lunar Laser Ranging \\citep[LLR;][]{williams04a}. Our VLBI observations and results are presented in \\S\\ref{sec:vlbi} and \\S\\ref{sec:results} respectively and the limitations on apparent accelerations (due to \\Gdot\\ or other possible causes such as unseen massive planets) are derived in \\S\\ref{sec:anomacc}. In \\S\\ref{sec:gwb}, we investigate the impact of the stochastic GWB on our results and derive an independent limit on the GWB amplitude. We summarize our conclusions in \\S\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} We have observed PSR J0437--4715 using the LBA and obtained the most precise pulsar distance determination to date, with an error $<1.5$\\,pc. Combined with accurate timing data, this has enabled us to confirm that $|\\Gdot/G| < 3.1 \\times 10^{-12}$\\ yr$^{-1}$, the most stringent limit not obtained though Solar System tests. Alternatively, assuming an unchanging gravitational force, the results can be interpreted as excluding any unseen Jupiter--mass planets within 226 AU of the Sun in 50\\% of the sky. Finally, the agreement between VLBI and timing results in this single case implies an upper limit to the stochastic GWB amplitude which is within an order of magnitude of the best limit derived from observations of ensembles of pulsars." }, "0808/0808.3628_arXiv.txt": { "abstract": "We investigate the direct contribution of strong, sunspot-like magnetic fields to helioseismic wave travel-time shifts via two numerical forward models, a 3D ideal MHD solver and MHD ray theory. The simulated data cubes are analyzed using the traditional time-distance center-to-annulus measurement technique. We also isolate and analyze the direct contribution from purely thermal perturbations to the observed travel-time shifts, confirming some existing ideas and bring forth new ones: (i) that the observed travel-time shifts in the vicinity of sunspots are largely governed by MHD physics, (ii) the travel-time shifts are sensitively dependent on frequency and phase-speed filter parameters and the background power below the $p_1$ ridge, and finally, (iii) despite its seeming limitations, ray theory succeeds in capturing the essence of the travel-time variations as derived from the MHD simulations. ", "introduction": "Local helioseismic diagnostic methods such as time-distance helioseismology \\citep{duvalletal1993}, helioseismic holography \\citep{lb1997} and ring-diagram analysis \\citep{hill1988}, have over the years provided us with unprecedented views of the structures and flows under sunspots and active regions. However, a growing body of evidence appears to suggest that interpretations of the measured statistical changes in the properties of the wave-field may be rendered inaccurate by complexities associated with the observations and wave propagation physics. Incorporating the full MHD physics and understanding the contributions of phase and frequency filters, and differences in the line formation height, are thought to be central to future models of sunspots. One of the earliest studies that highlighted the interaction of waves with sunspots was the Fourier-Hankel analysis of \\cite{bdl1987}, who found that sunspots can absorb up to half of the incident acoustic-wave power and shift the phases of interacting waves quite significantly (see also \\citealt{braun1995}). These results were echoed over the years by a steady steam of theoretical results (e.g. \\citealp{bogdanetal1996}; \\citealp{cb1997}; \\citealp{ccb2003}; \\citealp{crouchetal2005}; \\citealp{cally2007}) that have consistently emphasized the need for more sophisticated modeling and interpretation of wave propagation in strongly magnetized regions. Important advances in our observational understanding of sunspots were also achieved by \\cite{duvalletal1996} and \\cite{zkd2001}, who inferred the presence of flows underneath sunspots, and \\cite{kds2000} who estimated the sub-surface wave-speed topology. However, while the inversion procedures applied to derive these results fail to directly account for the tensorial nature of magnetic field effects, the action of the field is mimicked via changes in the acoustic properties of the medium (the so-called wave speed). Recently however, numerical forward models of helioseismic wave (e.g. \\citealp{cgd2008}; \\citealp{hanasoge2008}) and ray \\citep{mc2008} propagation in magnetized atmospheres have been developed and are beginning to make inroads into this problem. In particular, the results of \\cite{mc2008} and Cameron (2008; private communication) strongly suggest that active-region magnetic fields play a substantial role in influencing the wave field, and that the complex interaction of magnetic fields with solar oscillations, as opposed changes in the wave-speed, are the major causes of observed travel-time inhomogeneities in sunspots. We study the impact of strong magnetic fields on wave propagation and the consequences for time-distance helioseismology using two numerical forward models, a 3D ideal MHD solver and MHD ray theory. The simulated data cubes are analyzed using the traditional surface-focused center-to-annulus method frequently applied in the time-distance analyses of sunspots (e.g., \\citealp{cbk2006}). Furthermore, we apply the same method as \\cite{mc2008} to also isolate and analyze the thermal contribution to the observed travel-time shifts. ", "conclusions": "Incorporating the full MHD physics into the various forward models used in local helioseismology is essential for testing inferences made in regions of strong magnetic fields. By comparing numerical simulations of MHD wave-field and ray propagation in a model sunspot, we find that: i) the observed travel-time shifts in the vicinity of sunspots are strongly determined by MHD physics, although sub-surface thermal variations also appear to affect ray timings by modifying the acoustic cut-off frequency, ii) the time-distance travel-time shifts are strongly dependent on frequency, phase speed filter parameters and the background power below the $p_1$ ridge, and finally iii) MHD ray theory succeeds in capturing the essence of center-to-annulus travel-time variations as derived from the MHD simulations. The most unsettling aspect about this analysis is that despite using a background stratification that differs substantially from Model S \\nocite{cdetal1996} (Christensen-Dalsgaard et al. 1996) and a flux tube that clearly lacks a penumbra, the time shifts still look remarkably similar (at least qualitatively) to observational time-distance analyses of sunspots. Preliminary tests conducted with different sunspot models (e.g., different field configurations, peak field strengths etc.) have also provided similar results. Given the self-consistency of these results, as derived from both forward models, it could imply that we are pushing current techniques of local helioseismology to their very limits. It would appear that accurate inferences of the internal constitution of sunspots await a clever combination of forward modeling, observations, and a further development of techniques of statistical wave-field analysis." }, "0808/0808.3905_arXiv.txt": { "abstract": "With a self-similar magnetohydrodynamic (MHD) model of an exploding progenitor star and an outgoing rebound shock and with the thermal bremsstrahlung as the major radiation mechanism in X-ray bands, we reproduce the early X-ray light curve observed for the recent event of XRO 080109/SN 2008D association. The X-ray light curve consists of a fast rise, as the shock travels into the ``visible layer\" in the stellar envelope, and a subsequent power-law decay, as the plasma cools in a self-similar evolution. The observed spectral softening is naturally expected in our rebound MHD shock scenario. We propose to attribute the ``non-thermal spectrum\" observed to be a superposition of different thermal spectra produced at different layers of the stellar envelope. ", "introduction": "SN 2008D, the best type Ibc supernova detected so far, is preceded by a X-Ray Outburst (XRO) captured by SWIFT satellite on 2008 January 9, and this XRO is interpreted as a shock breakout of a Wolf-Rayet (WR) progenitor with a radius of $\\sim 10^{11}$ cm \\citep {Soderberg2008}. The isotropic X-ray energy is estimated to be $\\sim 2\\times10^{46}$ erg, and there seems no collimation detected so the event is not regarded as a GRB. This XRO showed a rapid rise, peaked at $\\sim 63$ s, and a decay modelled to be exponential with an e-folding time of $\\sim 129$ s \\cite{Soderberg2008}. The follow-up optical and ultraviolet observations indicate a total supernova kinetic energy of $\\sim 2-4\\times10^{51}$ erg and a mass of SN ejecta to be $\\sim 3-5$ M$_{\\odot}$ \\cite{Soderberg2008}. Some authors estimate from a detailed spectral analysis that SN 2008D, originally a $\\sim 30$ M$_{\\odot}$ star, has a spherical symmetric explosion energy of $\\sim 6\\times10^{51}$ erg and an ejected mass $\\sim 7$ M$_{\\odot}$ \\cite{Mazzali}. The evolution of optical spectra of XRO-SN 2008D resembles that of XRO-SN 2006aj, whose progenitor is also believed to be a WR star \\cite{Campana2006}. The production of $\\gamma$-rays and X-rays by shock breakouts has been proposed earlier \\citep{Colgate, Chevalier}. This XRO and the associated SN present an unprecedented case to be investigated in details, especially on interpretations for the rise and decay times of the X-ray light curve. The claim of an exponential decay may be premature given a fairly large scatter, and it may have concealed valuable physical clues offered by this XRO. During the XRO, the observed spectroscopic softening still lacks a convincing explanation. Here, we advance a self-similar MHD rebound shock model in an attempt to reproduce the observed X-ray light curve. The next section contains an overall description of the self-similar MHD model and the procedure of analysis; in the third section, we compare our model results with data; and conclusions are summed up in the last section. ", "conclusions": "With a self-similar MHD void shock model and the thermal bremsstrahlung as the main radiation loss, we obtain a fairly good fit to the X-ray light curve observed and confirm that XRO 080109 is most likely a shock breakout event. We identify that the decay in the X-ray light curve follows a power law (instead of an exponential law) in time since the core collapse and rebounce, which occurred $\\sim 552$ s before the observation of X-ray emissions. Meanwhile, the spectral softening is expected qualitatively. In this work, we use the most simplified radiation transfer presumption that the radiation produced in the `visible layer' can be totally observed. Actually, the optical depth varies with radius and should be treated in a more elaborate manner. Additionally, we presume that the boundaries of the ``visible layer\" $r_{in}$ and $r_{out}$ do not vary with time. Despite all these idealizations, our self-similar dynamic approach appears to be suitable to couple with the radiation process and to model X-ray outbursts in supernova as observed. Regarding the X-ray spectra observed, they cannot be fitted with a simple blackbody profile, and a nonthermal power law profile was suggested \\cite{Soderberg2008}. Several mechanisms have been proposed to explain such X-ray spectra, for example the bulk comptonization by scatterings of the photons between the ejecta and a dense circumstellar medium \\cite{Soderberg2008}, or diluted thermal spectra which require the thermalization occurs at a considerable depth in the supernova \\cite{Chevalier2008}. We propose that the power-law profile might be a natural result of multi-colour superposition of blackbody spectra. Based on our scenario outlined here, X-ray emissions come from different layers within the radiative layer around the stellar surface, and the radiative layer has different temperatures at different depth. As a result, the observed X-ray spectra are the superposition of thermal blackbody components with different temperatures. We suggest that this might resolve issues of spectral profile and evolution. During breakouts of rebound shocks and in the presence of MHD shock accelerated relativistic electrons usually presumed with a power-law energy spectrum within a certain electron energy range, we could also compute synchrotron emissions associated with such kind of SN shock breakouts. Among others, it is then possible to follow the evolution of magnetic field strength associated with SN explosions \\cite{Lou94, LouWang07} and estimate the effectiveness of accelerating relativistic particles (i.e., high-energy cosmic rays \\cite{Science06}). There is the freedom of choosing a few parameters to fit the data at a certain epoch. It is then possible to test the hypothesis of a self-similar shock evolution by further observations. In a more general perspective and on the basis of our dynamic models for rebound MHD shocks, we hope to further develop radiative diagnostics for shock breakouts of supernovae and thus for SN related GRBs. \\begin{theacknowledgments} This research has been partially supported by Tsinghua Center for Astrophysics (THCA), by the National Natural Science Foundation of China (NSFC) grants 10373009 and 10533020 and by the National Basic Science Talent Training Foundation (NSFC J0630317) at the Tsinghua University, and by the SRFDP 20050003088 and the Yangtze Endowment from the Ministry of Education at Tsinghua University. \\end{theacknowledgments}" }, "0808/0808.0137_arXiv.txt": { "abstract": "The concept of pyramid wavefront sensors (PWFS) has been around about a decade by now. However there is still a great lack of characterizing measurements that allow the best operation of such a system under real life conditions at an astronomical telescope.\\newline In this article we, therefore, investigate the behavior and robustness of the pyramid infrared wavefront sensor PYRAMIR mounted at the 3.5 m telescope at the Calar Alto Observatory under the influence of different error sources both intrinsic to the sensor, and arising in the preceding optical system. The intrinsic errors include diffraction effects on the pyramid edges and detector read out noise. \\newline The external imperfections consist of a Gaussian profile in the intensity distribution in the pupil plane during calibration, the effect of an optically resolved reference source, and noncommon-path aberrations. We investigated the effect of three differently sized reference sources on the calibration of the PWFS. For the noncommon-path aberrations the quality of the response of the system is quantified in terms of modal cross talk and aliasing. We investigate the special behavior of the system regarding tip-tilt control. \\newline From our measurements we derive the method to optimize the calibration procedure and the setup of a PWFS adaptive optics (AO) system. We also calculate the total wavefront error arising from aliasing, modal cross talk, measurement error, and fitting error in order to optimize the number of calibrated modes for on-sky operations. These measurements result in a prediction of on-sky performance for various conditions. ", "introduction": "\\label{sec:Intr} Within the context of AO systems pyramid wavefront sensors (PWFS) are a relatively novel concept. The special interest in this concept arises from the prediction of a gain in sensitivity -- and thus in limiting magnitude -- for a nonmodulated PWFS over a Shack-Hartmann sensor (SHS) \\cite{rag99} in closed-loop conditions for a well-corrected point source. The definition of this gain is that in order to achieve the same correction quality the PWFS needs less signal than the SHS. This gain in sensitivity results basically from the fact that the accuracy of the measurement and the resulting reconstruction error $\\sigma^2_{SH}$ of the SHS depends on $\\lambda\\over{d_{sub}}$ with $\\lambda$ = sensing wavelength, $d_{sub}$ = subaperture size typically chosen on the order of the Fried-parameter $r_0$ for typical site seeing at science wave-length. In the case of the PWFS the measurement error $\\sigma^2_{P}$ depends on $\\lambda\\over{D}$ with $D$ being the telescope diameter ($D >> r_0$). From these formulas the difference in stellar magnitudes that are needed to achieve the same Strehl ratio for the exemplary case of tip-tilt only can be derived for both sensors as $\\Delta m = -2.5 log\\left({\\sigma_P^2\\over{\\sigma_{SH}^2}}\\right)\\approx -2.5 log\\left({r_0^2\\over{D^2}}\\right)$. Simulations show , for instance, \\cite{rag99} that the gain in sensitivity for a 4\\,m class telescope and a seeing of $0.5''$ ($r_0=20$\\,cm) is predicted to be 2 magnitudes. \\subsection{The Pyramid Principle}\\label{subsec:PP} Wavefront sensing based on the pyramid principle has its origin in the Foucault knife-edge test. The historical development of the PWFS is described in \\cite{Camp2006}. \\begin{figure}[h!] \\centering \\includegraphics[width=12cm]{f1.jpg} \\caption{(Courtesy of S. Egner) The pyramid principle. An example of corresponding pixels is marked. The gradient in this position of the pupil is calculated from the intensity differences in between these pixels.} \\label{im:f1} \\end{figure} The optical setup of a pyramid sensor is shown in Figure \\ref{im:f1}. The transmissive, four-sided pyramid prism is placed in the focal plane. The focus is placed on the tip of the pyramid. After the pyramid, a relay lens images the pupils onto the detector.\\newline The signal a four-sided AO PWFS system uses are the intensities inside four pupil images. The illumination of these images depends on the aberrations of the wavefront. The signal $S$ one extracts is the difference in intensities $I_{1,2,3,4}$ between corresponding pixels in the four pupils ,i.e., the pixels at the same optical position in the pupils as shown in Fig.\\ref{im:f1}: \\begin{equation}S_x(x,y)={{I_1(x,y)+I_3(x,y)-[I_2(x,y)+I_4(x,y)]}\\over{I_1(x,y)+I_2(x,y)+I_3(x,y)+I_4(x,y)}}\\end{equation} \\begin{equation}S_y(x,y)={{I_1(x,y)+I_2(x,y)-[I_3(x,y)+I_4(x,y)]}\\over{I_1(x,y)+I_2(x,y)+I_3(x,y)+I_4(x,y)}}\\end{equation} for the x- and y-direction, respectively. These signals $S$ are what we will refer to as gradients, even if they might not exactly represent wavefront gradients. In the limit of small perturbations and a telescope with infinite aperture the frequency spectrum of the signal $S_x$ of our sensor is given by \\begin{equation}\\widetilde{S_x}=i{\\rm sgn}(u)\\widetilde{\\phi}(u,v).\\end{equation} Here $\\widetilde{()}$ means Fourier transform, $\\phi(u,v)$ the phase of the electromagnetic wave with $u$ and $v$ the coordinates in Fourier space, and sgn the sign-function (see \\cite{Costa2003}). Thus the sensor is working as a phase sensor in this regime. \\newline The principle of the PWFS implies some limitations that either do not occur in other sensor types, or have a much more \"dramatic\" impact on PWFSs than on other sensor types. In this class fall the structure of the pyramid edges that cause both diffraction and scattering, the read out noise of the system, the goodness of centering the beam on the pyramid tip, noncommon-path aberrations, and the homogeneity of the pupil illumination, the latter being important especially during calibration. It is these limitations, that may ultimately influence the choice of this or another sensor type, that will be examined in the course of this article. \\newline The structure in the remaining part of the article is as follows: In the next section, the PYRAMIR system is described in detail. The optical path and the possible detector read out modes are explained. Section 3 accounts for the calibration procedure of the system. The peculiarities of tip-tilt calibration and flattening the wavefront are presented. Section 4 shows fundamental limitations to the performance of a pyramid system. ``Fundamental'' in this case is not meant to be based on natural laws and constants, but on always-present aberrations and nonideal conditions. Thus, the fundamental limitations discussed here can be eased by careful alignment and set up of the system, but they can never entirely be removed. In this context, we explore the effect of static aberrations, different calibration light sources, and diffraction and scattering on the pyramid edges to the response of the system. In the next section the implications of the modal cross talk of the system aliasing and measurement error to the number of modes to be calibrated, and the residual wavefront error are calculated. We end the section with a prediction of the on-sky performance under different seeing conditions. The last section concludes the results of our measurements and the implications for any pyramid wavefront sensor.\\newline On-sky results will be presented in a following paper \\cite{Peter2008}. ", "conclusions": "\\label{sec:conclusion} In this article we presented laboratory measurements performed with the pyramid wavefront sensor PYRAMIR. After a short introduction into the history of PWFS and their working principles, we presented the PYRAMIR system as infrared PWFS and explained the details of the system especially the possible read out modes of the detector. This can reach a speed of about 300 Hz with a RON of 20 $e^-$ rms. The calibration procedure of the system was described in detail. The pitfalls of TT-calibration were discussed in detail leading to the conclusion that a static part in TT will reduce the limiting magnitude and enhance TT-jitter and cannot be tolerated. Therefore we included a static TT-part in the bias pattern of the deformable mirror (dmBias) to perfectly center the beam. Also it has to be guaranteed that the amplitudes of calibration are the same in both axis in order to gain similar performance in both directions.\\newline Some of the fundamental sources of reduced performance were examined. We found that the amount of light diffracted out of the pupil images is 50\\% for a flat wavefront decreasing nonlinearly up to 2 $rad$ wavefront error where it becomes almost stable at 20\\%. A comparison with simulations shows that in our case this loss results from diffraction at the pyramid edges and imperfections in the optics. The latter outweighing for aberrations stronger than about 1.5 $rad$.\\newline The effects of a Gaussian illumination of the pupils, as well as the effect of an extended calibration light source, were investigated. Both have strong influence on the sensitivity of the system and should be avoided. On the other hand an extended target during the measurement does only slightly effect the performance. Thus the calibration light source should be as point like as possible for all applications.\\newline The effect of RON on the limiting magnitude of the guide star has been widely discussed. In the case of infrared detectors this noise is quite high. In our case 20 $e^-$ rms per pixel. This reduces the limiting magnitude by 5 mag. We investigated into the best possible mode set between the eigen modes of PYRAMIR, of the DM and the KL modes. The last two mode sets seem promising for the use on-sky. The latter has a larger linear regime than the eigen modes of the DM. The difference is small and, therefore, the correction error will not vary by much. Still the larger linear regime might help to close the loop under bad seeing conditions. \\newline We tested the best treatment of noncommon-path aberrations by applying artificial modes to the DM. The best treatment turned out to direct these aberrations completely into the path of the sensor. A possible better way using two calibrations, one for modes with positive amplitude one for those with negative amplitude was proposed to reduce the error of the mode with static aberrations, but put aside due to the fact that it was not running stably during the testing on-sky.\\newline The importance of a small calibration amplitude to minimize the resulting reconstruction error was shown. \\newline After the measurement of these fundamental properties, the subject of the best number of modes to calibrate was addressed. To solve this we investigated into the behavior of modal cross talk, aliasing, and measurement error dependence on the number of modes calibrated. We found that the averaged aliasing coefficient varies only slightly with the number of modes whereas the average modal cross talk coefficient decreases inversely linear with this number and the measurement error rises linearly with this number. Including the (theoretical) contribution of the KL modes in the wavefront error on sky we could show that the contribution of aliasing rises like $N^{(2-\\sqrt{3})\\over 2}$; the contribution of cross talk becomes constant for larger $N$ and the measurement error rises linearly with $N$. In CL only the error due to modal cross talk changes with respect to the open loop measurement. It will decrease because the modes that are corrected will contribute less to the modal cross talk than in open loop. The error of the residual wavefront decreases like $N^{-{\\sqrt{3}\\over 2}}$. Altogether we could show that at the border of the limiting magnitude the fitting error surpasses all other errors for a low number of corrected modes but will be overpowered by the measurement error at about the place of the optimum number of modes to be corrected. For the PYRAMIR system this is 5 modes. The other errors will become important for even slightly brighter stars. Here cross talk and aliasing error are almost identical in strength. In the case of noncommon-path aberrations in the system the modal cross talk and aliasing errors will rise. Modal cross talk increases linearly, aliasing error stays constant until the static aberration reaches the border of the linear regime of the sensor, then it increases nonlinearly. \\newline From the entire error-budged we could predict the performance on-sky for various seeing conditions. For a seeing of 1'' and a wind speed on ground level of 5$ms^{-1}$ we can achieve a 39\\% Strehl ratio in $K^\\prime$-band, but we have to run at about maximum frame rate (300 Hz). The limiting S/N per subaperture on the detector is about 0.25 or 6.0 mag in $K^\\prime$. For good seeing conditions and a moderate loop band width, the predicted Strehl ratio will be about 60\\% for bright stars. Again the limiting S/No per subaperture on the detector is about 0.25 or 7.2 mag in $K^\\prime$." }, "0808/0808.2074_arXiv.txt": { "abstract": "Due to the high efficiency of planet detections, current microlensing planet searches focus on high-magnification events. High-magnification events are sensitive to remote binary companions as well and thus a sample of wide-separation binaries are expected to be collected as a byproduct. In this paper, we show that characterizing binaries for a portion of this sample will be difficult due to the degeneracy of the binary-lensing parameters. This degeneracy arises because the perturbation induced by the binary companion is well approximated by the Chang-Refsdal lensing for binaries with separations greater than a certain limit. For binaries composed of equal mass lenses, we find that the lens binarity can be noticed up to the separations of $\\sim 60$ times of the Einstein radius corresponding to the mass of each lens. Among these binaries, however, we find that the lensing parameters can be determined only for a portion of binaries with separations less than $\\sim 20$ times of the Einstein radius. ", "introduction": "Searches for extrasolar planets by using microlensing are being carried out by observing stars located toward the Galactic bulge field. The lensing signal of a planet is a short-duration perturbation to the smooth standard light curve of the primary-induced lensing event occurring on a background star \\citep{mao91, gould92}. For the detections of the short-duration planetary lensing signals, these searches are using a combination of survey and follow-up observations, where the survey observations (e.g., OGLE, \\citet{udalski03}; MOA, \\citet{bond02}) aim to maximize the lensing event rate by monitoring a large area of sky and the follow-up observations (e.g., PLANET, \\citet{albrow01}; MicroFUN, \\citet{dong06}) intensively monitor the events alerted by the survey observations. However, the limited number of telescopes restricts the number of events that can be followed at any given time and thus priority is given to those events that will maximize the planetary detection probability. Currently, the highest priority is given to high-magnification events because the source trajectories of these events always pass close to the perturbation region around the planet-induced caustic located near the primary lens \\citep{griest98}. In addition, follow-up observations can be prepared for these events because the perturbation typically occurs near the peak of the event, which can be predicted from the data on the rising part of the event light curve. As a result, six (OGLE-2005-BLG-071Lb, OGLE-2005-BLG-169Lb, OGLE-2006-BLG-109Lb,c, MOA-2007-BLG-400b, MOA-2007-BLG-192) of the eight reported microlensing planets were detected through the channel of high-magnification events \\citep{bond04, udalski05, beaulieu06, gould06, gaudi08, dong08, bennett08}. In addition to planets, high-magnification events are sensitive to wide-separation binary companions as well. Similar to the planetary case, the companion of a wide-separation binary induces a small caustic close to the primary lens and thus can produce a short-duration perturbation near the peak of a high-magnification event. Due to the different nature of the companions, however, the perturbations induced by a wide-separation binary companion and a planet can be distinguished \\citep{albrow02, han08}. Therefore, under the current planetary lensing strategy focusing on high-magnification events, a considerable number of wide-separation binaries are expected to be detected as a byproduct and this sample might provide useful information about the physical distribution of binaries. Unlike this expectation, however, we find that characterizing binaries for a significant portion of the binary lens sample will be difficult due to the degeneracy of the binary-lensing parameters. This degeneracy arises because the perturbation induced by the binary companion is well approximated by the Chang-Refsdal lensing (hereafter C-R lensing) for binaries with separations greater than a certain value. The C-R lensing represents single-mass lensing superposed on a uniform background shear. For a wide-separation binary, the shear results from the combination of the binary-lens parameters and thus the individual parameters cannot be separately determined. The paper is organized as follows. In \\S\\ 2, we briefly describe the properties of binary and C-R lensing. In \\S\\ 3, we demonstrate the proximity between the binary and C-R lensing for binaries with separations beyond a certain limit. We then set the range in the binary-lensing parameter space where the degeneracy of binary-lensing parameters occurs. We discuss the meaning of this degeneracy in the studies of binaries by using microlensing. We summarize the results and conclude in \\S\\ 4. ", "conclusions": "By investigating the patterns of central perturbations produced by wide-separation binary companions, we found that the perturbation can be noticed and thus lens binarity can be identified for binaries with considerable separations. However, we also found that perturbations for a significant portion of these binaries can be well approximated by the C-R lensing, implying that the individual binary-lensing parameters cannot be separately determined and thus characterization of the binary is difficult. We set the range in the binary-lensing parameter space where this degeneracy occurs. From this, we found that the lens binarity can be noticed up to the separations of $\\sim 60$ times of the Einstein radius corresponding to the mass of each lens of a binary composed of equal mass lenses. Among these binaries, however, we found that the lensing parameters can be determined only for a portion of binaries with separations less than $\\sim 20$ times of the Einstein radius." }, "0808/0808.0889_arXiv.txt": { "abstract": "We report the detection of very high-energy $\\gamma$-ray emission from the intermediate-frequency-peaked BL\\,Lacertae object W\\,Comae ($z = 0.102$) by VERITAS, an array of four imaging atmospheric-Cherenkov telescopes. The source was observed between January and April 2008. A strong outburst of $\\gamma$-ray emission was measured in the middle of March, lasting for only four days. The energy spectrum measured during the two highest flare nights is fit by a power-law and is found to be very steep, with a differential photon spectral index of $\\Gamma = 3.81 \\pm 0.35_{\\rm{stat}} \\pm 0.34_{\\rm{syst}}$. The integral photon flux above $200 \\, \\rm{GeV}$ during those two nights corresponds to roughly $9\\%$ of the flux from the Crab Nebula. Quasi-simultaneous Swift observations at X-ray energies were triggered by the VERITAS observations. The spectral energy distribution of the flare data can be described by synchrotron-self-Compton (SSC) or external-Compton (EC) leptonic jet models, with the latter offering a more natural set of parameters to fit the data. ", "introduction": "The blazars detected at very high energies (VHE, $E > 100 \\, \\rm{GeV}$) by ground-based imaging atmospheric-Cherenkov telescopes (IACTs) are extreme objects in the active galactic nuclei (AGN) population. Typically these sources show core-dominated emission, and they are characterized by rapid variability and strong broadband continuum emission ranging from the radio band to the X-ray band. Multi-wavelength data on blazars reveal that their spectral energy distribution (SED) is characterized by two broad, well-separated ``humps'' arising from synchrotron (low-energy) and inverse-Compton (IC) or hadronic emission (high-energy). Blazars are categorized into different sub-classes based on the frequencies at which these emission components reach a maximum. Flat-spectrum radio quasars (FSRQs) and low-frequency-peaked BL\\,Lacs (LBLs) are generally seen to have low-frequency, synchrotron peaks in the IR/optical regime, whereas high-frequency-peaked BL\\,Lacs (HBLs) exhibit peaks in the X-ray band, in several cases at energies of $\\sim$100~keV or higher. Intermediate-frequency-peaked BL\\,Lacs (IBLs) bridge the gap between LBLs and HBLs. The properties of the broad sub-classes of blazars, the luminosity-versus-frequency trends and possible physical explanations are discussed by \\cite{ghi08}. Gamma rays are an important component of the SED of blazars; the integral power in the $\\gamma$-ray waveband is comparable or higher than that in the rest of the electromagnetic spectrum (from radio to X-rays). There are $65$ blazars detected at MeV/GeV energies by the EGRET instrument on board the {\\sl Compton Gamma Ray Observatory (CGRO)} \\citep{mat01, har99} and several of the other EGRET sources also have likely blazar counterparts. Ground-based IACTs have established $\\sim$20 blazars as emitters of VHE $\\gamma$-radiation; see for example \\citet{wak07}. While the blazars detected at MeV/GeV energies tend to be largely FSRQs (and some LBLs), almost all VHE blazars belong to the class of HBLs, the only exceptions being the LBLs BL\\,Lacertae \\citep{alb07}, S5\\,0716+71 \\citep{tes08} and the FSRQ 3C\\,279 \\citep{tes07}. The IBL W\\,Comae (W\\,Com) at a redshift of $z = 0.102$ has long been an object of interest for VHE observatories. W\\,Com was discovered at radio frequencies \\citep{bir71} and later detected at X-ray energies by the {\\sl Einstein} Imaging Proportional Counter in June 1980 \\citep{wor90}. Data taken with the {\\sl BeppoSAX} satellite in 1998 \\citep{tag00} clearly showed that the transition between the low- and high-energy peaks in the SED occurs around $\\sim$4~keV. In April-May 1998, an exceptional optical outburst was detected from W\\,Com showing rapid variability on time scales of hours \\citep{mas99}. At $\\gamma$-ray energies, W\\,Com was detected by EGRET in the $100 \\, \\rm{MeV} - 10 \\, \\rm{GeV}$ band \\citep{har99} and in a re-analysis of the data up to $25 \\, \\rm{GeV}$ \\citep{din01}. Due to its rather hard EGRET spectrum (photon spectral index $\\alpha=1.73\\pm 0.18$), with no sign of spectral cut-off \\citep{har99}, the source became even more interesting for VHE observations. However, W\\,Com was not detected by the Whipple IACT above $300 \\, \\rm{GeV}$ in 1993/94 \\citep{ker95} and 1995/96/98 \\citep{hor04}, nor by the solar heliostat Cherenkov telescope STACEE \\citep{sca04}. In this paper we report the discovery of VHE $\\gamma$-ray emission from W\\,Com with VERITAS. ", "conclusions": "VERITAS detected VHE $\\gamma$-ray emission from W\\,Com with a statistical significance of $4.9$~standard deviations for the entire data set (January to April 2008). A strong outburst was observed in March 2008 with a statistical significance of $>8$~standard deviations, that lasted for only four days. In addition to W\\,Com, a second extragalctic source (the VHE blazar 1ES\\,1218+304) is detected in the same field of view -- for the first time in VHE $\\gamma$-ray astronomy. W\\,Com is the first VHE-detected blazar of the IBL class. The extension of the VHE catalog to the FRSQ, LBL and IBL classes will play a major role in our understanding of blazar populations and dynamics. The quasi-simultaneous SED of W\\,Com at the time of the VHE outburst can be modeled with a simple one-zone SSC model. However, an unusually low magnetic field of $B = 0.007 \\, \\rm{G}$ (more than an order of magnitude lower than typically found in the modeling of other BL\\,Lac-type blazars) and a small ratio of the magnetic field to electron energy density of $\\sigma = 1.3 \\cdot 10^{-3}$ are required. An EC model with more natural parameters ($B=0.36 \\, \\rm{G}$ and $\\sigma = 1$) provides a good fit and could account for shorter variability time scales. Our model results agree with the expectation that for IBLs (and LBLs) the higher optical luminosity plays an important role in providing the seed population for IC scattering. The IBL W\\,Com will be an excellent target for future observations at GeV energies with GLAST and in the VHE regime with IACTs, including correlated GeV/TeV variability studies." }, "0808/0808.2248_arXiv.txt": { "abstract": "(1270) Datura is the largest member of a very young asteroid cluster that was thought to be broken-up 0.45 Myr ago. The light-curve and the rotation-resolved reflectance spectra (0.6 $\\mu$m -- 1.0 $\\mu$m) were observed in order to find ``fresh'' surface. Our data show no significant spectral variation along the rotation phase. The depth of the 0.95 $\\mu$m absorption band, which indicates the degree of space weathering, was similar to that of an old S-type asteroid. This suggests that the reflectance spectrum in this wavelength range changes rapidly and saturates the depth of the 0.95 $\\mu$m absorption in less than 0.45 Myr in the main belt environment. ", "introduction": "A parent body of an ordinary chondrite, one of the most common stony meteorites, is believed to be a S-type asteroid that is commonly distributed in the inner main belt region and is thought to consist of silicate rocks covered with regolith. However, there are discrepancies between the reflectance spectra of ordinary chondrites and S-type asteroids: S-type asteroids show redder spectral slope in the visible wavelength and the depth of the absorption band at $\\sim0.95$ $\\mu$m is shallower than that of ordinary chondrites \\citep{cha73}. These spectral discrepancies are attributed to ``space weathering,'' which alter a fresh, chondritic spectra to the S-type asteroid spectra (e.g., Clark et al. 2002). Although details of the physical process of the space weathering have not been fully understood, it is thought that irradiation of high energy particles (solar wind or cosmic ray) and/or bombardment of micro-meteorites makes nanophase iron in the surface of regolith that cases the spectral change \\citep{ada71, hap00, pie00, sas01}. One of the unsolved issues on space weathering is to determine the time scale of the weathering process. Dynamical investigations of asteroid families are enabling us to create a time scale of the history of asteroid evolution (e.g., Nesvorn\\'y et al. 2005). \\citet{nes05} related the family age and the spectral slope, and \\citet{wil08} revised their results and found that the time scale of the spectral slope change by space weathering was $570\\pm220$ Myr. However, \\citet{cha07} and \\citet{ver07} observed (832) Karin, whose age was estimated to be 5.8 Myr, and found that the depth of 0.95 $\\mu$m absorption band was as shallow as a typical S-type asteroid. These observations suggest that the depth of 0.95 $\\mu$m absorption band changes faster than the reddening of the spectral slope. Recently \\citet{nes06a} found a very young asteroid cluster, Datura cluster, whose age was estimated to be $0.45\\pm0.05$ Myr. Datura cluster has been exposed for only $1/10$ of the age of the Karin cluster. Thus this cluster is a good target to examine the alteration speed by space weathering at the 0.95 $\\mu$m absorption band. (1270) Datura has a diameter of about 11 km and is orbiting in the inner main belt region ($a=2.2$ AU). I have observed (1270) Datura, which is the largest member of the Datura cluster, expecting to find a ``fresh'' region on its surface because the space-weathered old surface was removed by the cluster-forming break-up and a fragment might expose a fresh, interior surface on the whole area or at least in some area. In this letter I will report the rotation-resolved reflectance spectra of (1270) Datura and its implication for the space weathering. ", "conclusions": "The rotation curve shows fairly large amplitude that means we are not seeing (1270) Datura from above its pole, thus we see most of its surface when observing nearly a full rotational period. The fact that there is no significant difference among the spectra obtained at different rotation phases indicates that the surface of (1270) Datura has no significant large scale spatial variation in composition and in the degree of space weathering. The rotation speed of (1270) Datura is slow enough, and the size is large enough to be covered with regolith that must be newly created at the cluster-forming collision at 0.45 Myr ago and then deposited onto the surface. Since the deposition time scale might be longer compared to the rotation period of (1270) Datura, the dusts/gravels covered evenly over the surface. Thus it is natural that the surface of (1270) Datura is almost homogeneous when observed in a large spatial scale as discussed by \\citet{nes06b} and \\citet{ver07} for (832) Karin. To see the degree of space weathering of the surface of (1270) Datura, the phase-averaged spectrum is compared with that of an old S-type asteroid (15) Eunomia (estimated age is 2.5 Gyr \\citep{nes05}) and some of the ordinary chondrites as shown in Figure~\\ref{fig_comp}. Contrary to our expectation, the spectrum of (1270) Datura does not resemble those of ordinary chondrites, but is close to an old S-type asteroid (15) Eunomia. This means that the surface of (1270) Datura was already altered in the same degree with an old S-type asteroid in respect to the depth of 0.95 $\\mu$m absorption band. Because the surface was renewed 0.45 Myr ago by a cluster-forming collision, the change of the reflectance spectrum (from ordinary-chondrite-like one to that of S-type asteroids) must be completed in less than 0.45 Myr, if the age estimate by \\citet{nes06a} and the concept of space weathering are correct. \\citet{bin96, bin01} found that many of the near-earth objects (NEOs) are Sq- or Q-type, which connects S-type spectra to ordinary chondrite spectra. If NEOs have been space weathered in the same manner as (1270) Datura, the surface of those objects have been exposed much less than $0.45$ Myr. If the surface was exposed by a recent collision in the main belt, we can estimate the current collision rate in the main belt. However, we must be careful that the environment of space weathering in NEO region differs from the inner main belt. Also the establishment and stability of regolith might be different since the size of the objects of \\citet{bin96, bin01} are one order-of-magnitude smaller than (1270) Datura." }, "0808/0808.0417_arXiv.txt": { "abstract": "{In this contribution we discuss some of the main problems in high energy astrophysics, and the perspectives to solve them using different types of ``messengers'': cosmic rays, photons and neutrinos.} \\normalsize\\baselineskip=15pt ", "introduction": "The birth of high energy astrophysics can be traced to nearly a century ago, when the balloon flights of Victor Hess established that a form of ionizing radiation, that was soon given the name of ``cosmic rays'', was arriving from outer space. Soon it was demonstrated that these ``rays'' are in fact ultrarelativistic charged particles, mostly protons and fully ionized nuclei\\footnote{ Electrons have a steeper energy spectrum and constitute a fraction of only few percent of the flux; small quantities of positrons and anti--protons are also present.}, with a spectrum that extends to extraordinary high energies. The existence of a large flux of ultrarelativistic particles arrived completely unexpected, and its origin remained a mistery that only now is beginning to be clarified. The main reason for this very long delay in developing an understanding of this important physical phenomenon is that cosmic rays (CR) do not point to their production sites, because their trajectory is bent by the magnetic fields that permeates both interstellar and intergalactic space. Today we know that our universe contains several classes of astrophysical objects where non--thermal processes are capable to accelerate charged particles, both leptons ($e^\\mp$) or hadrons (protons and nuclei), to very high energy. These relativistic particles, interacting inside or near their sources, can produce photons, and (in case of hadrons) neutrinos that then travel in straight lines allowing the imaging of the sources. At the highest energy also the magnetic bending of charged particles can become sufficiently small to allow source imaging. A detailed study of the ``high energy universe'' can therefore in principle be performed using three different ``messengers'': photons, neutrinos and the cosmic ray themselves. This ``multi--messenger'' approach is still in its infancy, but has the potential to give us deep insights about the sites and physical mechanisms that produce these very high energy particles. Many of the proposed (or detected) acceleration sites are also associated with the violent acceleration of large macroscopic masses (one example is Gamma Ray Bursts), and therefore a fourth ``messenger'': gravitational waves has the potential to give very important and complementary information about these astrophysical environments. ", "conclusions": "" }, "0808/0808.4002_arXiv.txt": { "abstract": "The main aim of this study is the comparison of gravitational waveforms obtained from numerical simulations which employ different numerical evolution approaches and different wave-extraction techniques. For this purpose, we evolve an oscillating, nonrotating, polytropic neutron-star model with two different approaches: a full nonlinear relativistic simulation (in three dimensions) and a linear simulation based on perturbation theory. The extraction of the gravitational-wave signal is performed via three methods: the gauge-invariant {\\it curvature-perturbation} theory based on the Newman-Penrose scalar $\\psi_4$; the gauge-invariant Regge-Wheeler-Zerilli-Moncrief {\\it metric-perturbation} theory of a Schwarzschild space-time; some generalization of the quadrupole emission formula. ", "introduction": "\\label{sec:intro} The computation of the gravitational-wave emission from compact sources like supernova explosions, neutron-star oscillations and the inspiral and merger of two compact objects (like neutron stars or black holes) is one of the most lively subjects of current research in gravitational-wave astrophysics. This goal may be pursued using different numerical approaches. That is, (i) solving the {\\it full set} of coupled Einstein and matter equations; (ii) solving the {\\it linearized} Einstein and matter equations around a fixed background, when such an approximation is valid. In the latter case, with the additional condition of spherical symmetry, the formalism we employ is based on a multipolar expansion and the computation of the gravitational waves directly follows from the knowledge of the perturbative metric multipoles $k_\\lm$, $\\chi_\\lm$ and $\\psi_\\lm$. On the other hand, extracting gravitational waveforms from a space-time computed numerically in a given coordinate system is a highly nontrivial problem that has been addressed in various ways in the literature. In general, two routes have proven successful: (i) the gauge-invariant {\\it curvature-perturbation} theory based on the Newman-Penrose~\\cite{NP62} scalar $\\psi_4$, and (ii) the Regge and Wheeler~\\cite{RW57}, Zerilli~\\cite{Zerilli:1970se} theory of {\\it metric-perturbations} of a Schwarzschild space-time, recast in a gauge-invariant framework following the work of Moncrief~\\cite{Moncrief:1974am}. The aim of our study is the computation of the gravitational waveforms emitted by the very controlled system constituted by a nonrotating polytropic relativistic star that oscillates nonisotropically around its spherically symmetric equilibrium configuration because of an axisymmetric perturbation. Our aim is to follow two (complementary) calculation procedures. On one hand, we perform a full 3+1 numerical simulation of the system, \\ie we compute a numerical solution of the Einstein equations without approximations except those of the numerical method itself. Because of its generality, this approach allows us to analyze different physical regimes, in particular, the case in which the ``perturbation'' is not small and nonlinear effects can play a relevant role with important consequences on the waveforms. On the other hand, we follow a perturbative approach based on the assumption that the perturbation is ``small''. If this is the case, one can (i) expand the metric around a fixed background (\\ie the Tolman-Oppenheimer-Volkoff solution), (ii) retain only the linear term of this expansion and (iii) solve the {\\it linearized} Einstein equations. In addition, since the star is nonrotating, one can factorize the angular dependence by means of a spherical-harmonic decomposition of the metric and matter fields, and, thus, only a 1+1 system of partial differential equations must be solved. The present work has much in common with Refs.~\\cite{Shibata:2003aw,Pazos:2006kz}, where a comparison of different extraction techniques has been performed. Following the same inspiration of Ref.~\\cite{Pazos:2006kz}, we exploit perturbative computations to obtain ``exact'' waveforms to compare with the numerical-relativity--generated ones. As done in Ref.~\\cite{Shibata:2003aw}, we use an oscillating neutron star as a test-bed system, but we consider a wider range of possible wave-extraction techniques. Since there is a copious literature dealing with the problem of gravitational-wave extraction in numerical relativity, we prefer not to mention here the main bibliographic references, but rather to address the reader to the references in Refs.~\\cite{Shibata:2003aw,Pazos:2006kz} and to the citations in the following text. The article is organized as follows. In Sec.~\\ref{sec:nsetp} we describe the numerical time-evolution methods and the gravitational-wave extraction techniques adopted. In Sec.~\\ref{sbsc:init} we introduce our choice of initial data and Sec.~\\ref{sec:res} is devoted to the presentation of our results. Conclusions that can be drawn from our results are discussed in Sec.~\\ref{sec:concl}. Standard dimensionless units $c=G=M_\\odot=1$ and a spacelike signature $(-,+,+,+)$ are used. Greek indices are taken to run from $0$ to $3$, Latin indices from $1$ to $3$ and we adopt the standard convention for the summation over repeated indices. ", "conclusions": "\\label{sec:concl} We have compared various gravitational-wave--extraction methods that are nowadays very popular in numerical-relativity simulations: (i) the Abrahams-Price~\\cite{Abrahams:1995gn} technique based on the gauge-invariant Regge-Wheeler-Zerilli-Moncrief perturbation theory of a Schwarzschild space-time; (ii) the extraction method based on Weyl curvature scalars, notably the $\\psi_4$ function; (iii) some (variations of) quadrupole-type formulas. We have applied these methods to extract gravitational radiation from 3D numerical-relativity simulations of the very controlled system represented by a neutron star (with polytropic EOS), that is oscillating nonradially due to an initial pressure perturbation. The simulations have been performed via the {\\tt Cactus-CCATIE-Carpet-Whisky} general-relativistic nonlinear code. This code evolves the full set of Einstein equations in full generality in the three spatial dimensions. The accuracy of the waveforms extracted from the simulations, using the three methods recalled above, has been assessed (for small perturbations) via a comparison with waveforms (assumed to be exact) computed by means of the {\\tt PerBACCo} perturbative code. This code is designed to evolve, in the time domain, the Einstein equations linearized around a TOV background. It is 1+1-dimensional (\\ie one temporal and one spatial dimension) and adopts a constrained evolution scheme. This latter choice allows for the computation of very long and very accurate time series and similarly accurate waveforms. The initial pressure perturbation $\\delta p$ is given as an ``approximate'' eigenfunction of the star, whose maximum is a fraction of the central TOV pressure $p_c$. We focused only on $\\l=2$, $m=0$, quadrupolar deformations, but we analyzed four values of the perturbation in order to cover the transition from the linear to the nonlinear oscillatory regimes. We have first presented results of simulations done using only the 1D {\\tt PerBACCo} code to assess the accuracy of our exact waveforms. We have performed very long (about 1~s) and accurate simulations to extract both mode frequencies and damping times. We have analyzed finite-radius effects, finding that observers should be placed at extraction radius $r>200M$ in order to have amplitude errors below 1.6\\%. In doing 3D simulations in the perturbative regime, ($10^{-3}\\lesssim {\\rm max}(\\delta p/p_c)\\lesssim 10^{-2}$), we have found that {\\it both} metric and curvature wave-extraction techniques generate waveforms that are consistent, both in amplitude and phasing, with the perturbative results. Each method, however, was found to have drawbacks. On one hand, the Zerilli-Moncrief function presents an unphysical burst in the early part of the waveform; on the other hand, the $\\psi_4$ scalar requires a polynomial correction to obtain the corresponding metric multipole. Our conclusion is that, in our setup, one needs both extraction methods to end up with accurate waveforms. For larger values of the initial perturbation amplitude, nonlinear effects in the 3D general-relativistic simulations are clearly present. The effective relative amplitude of the main modes of the extracted gravitational wave is smaller for larger amplitudes of the initial perturbation, because of mode couplings. The Fourier spectra of the rest-mass--density projections [see Eq.~\\eqref{eq:rholm}] highlight that couplings between radial and quadrupolar fluid modes are present. Our study represents the first confirmation, in fully general-relativistic simulations, of the results of Ref.~\\cite{Passamonti:2007tm}, obtained via a perturbative approach. In addition, we have shown that the (non-gauge-invariant) generalizations of the standard Newtonian quadrupole formula that we have considered can be useful tools to obtain accurate estimates of the frequency of oscillation. By contrast, amplitudes are always significantly under/overestimated, consistently with precedent observations of Refs.~\\cite{Shibata:2003aw,Nagar:2005cj}. Finally, we discussed in detail some systematic errors that occur in the early part of the waveform extracted \\'a la Abrahams-Price. These errors show up, in the early part of the Zerilli-Moncrief function, in the form of a burst of {\\it junk} radiation whose amplitude grows linearly with the extraction radius. We have proposed some heuristic explanation of this fact and reproduced a similar behavior in low-accuracy perturbative simulations. Globally, our conclusion is that the extraction of the Zerilli-Moncrief function from a numerical-relativity simulation can be a delicate issue: small errors can conspire to give totally nonsensical results. Typically, these errors will show up as parts of the waveform whose amplitude grows with the observer's radius. We have also implemented the generalized wave-extraction approach based on the formalism of Refs.~\\cite{Sarbach:2001qq,Martel:2005ir,Pazos:2006kz,Korobkin:2008ji}, without any evident benefit. Note, however, that these kind of problems encountered with the Abrahams-Price wave-extraction procedure (as well as with its generalized version) seem to appear {\\it specifically} in the presence of matter. In binary black-hole coalescence simulations curvature and metric waveforms seem to be fully consistent~\\cite{Pollney:2007ss}. This last remark leads us to suggest that the Abrahams-Price wave-extraction technique, a ``standardized'' and very basic procedure and infrastructure that has been developed long ago (and tested at the time) for specific applications to black-hole physics, should be rethought and reanalyzed when the Einstein equations are coupled to matter. For this reason, in the presence of matter, since systematic errors could be hard to detect and are present already in the simplest cases, we strongly encourage the community to make use of {\\it both} wave-extraction techniques (curvature as well as metric perturbations) and to be always prepared to expect inaccuracies in the metric waveforms. In addition, concerning the many advantages related to extracting the metric waveforms directly from the space-time, we believe that it is also urgent and important for the community to have reliable implementations of the Abrahams-Price technique based on the Sarbach-Tiglio-Martel-Poisson~\\cite{Sarbach:2001qq,Martel:2005ir,Pazos:2006kz,Korobkin:2008ji} formalism." }, "0808/0808.3414_arXiv.txt": { "abstract": "A set of 41 nearby stars (closer than 25 pc) is investigated which have very wide binary and common proper motion (CPM) companions at projected separations between $1000$ and $200\\,000$ AU. These companions are identified by astrometric positions and proper motions from the NOMAD catalog. Based mainly on measures of chromospheric and X-ray activity, age estimation is obtained for most of 85 identified companions. Color -- absolute magnitude diagrams are constructed to test if CPM companions are physically related to the primary nearby stars and have the same age. Our carefully selected sample includes three remote white dwarf companions to main sequence stars and two systems (55 Cnc and GJ 777A) of multiple planets and distant stellar companions. Ten new CPM companions, including three of extreme separations, are found. Multiple hierarchical systems are abundant; more than 25\\% of CPM components are spectroscopic or astrometric binaries or multiples themselves. Two new astrometric binaries are discovered among nearby CPM companions, GJ 264 and HIP 59000 and preliminary orbital solutions are presented. The Hyades kinematic group (or stream) is presented broadly in the sample, but we find few possible thick disk objects and none halo stars. It follows from our investigation that moderately young (age $\\lesssim 1$ Gyr) thin disk dwarfs are the dominating species in the near CPM systems, in general agreement with the premises of the dynamical survival paradigm. Some of the multiple stellar systems with remote CPM companions probably undergo the dynamical evolution on non-coplanar orbits, known as the Kozai cycle. ", "introduction": "\\label{firstpage} Components of wide stellar binaries and common proper motion pairs have been drawing considerable interest for many years. Despite the increasing accuracy of observations and the growing range of accessible wavelengths, the origin of very wide, weakly bound or unbound systems remains an open issue. \\citet{lep7} estimated that at least 9.5\\% of stars within 100 pc have companions with projected separations greater than 1000 AU. The renewed interest was boosted by the detection of a dearth of substellar mass companions in spectroscopic binaries, and by the attempts to account for the missing late-type members of the near solar neighborhood. The main objective of this paper is to investigate a well-defined set of nearby stars in very wide CPM pairs and to discover new pairs, possibly with low-mass companions. The secondary goal of our investigation is to establish or estimate the age and evolutionary status of bona fide companions using a wide range of available astrometric and astrophysical data. The origin and status of wide CPM systems is still a mystery, because most of them are likely unbound or very weakly bound and are expected to be easily disrupted in dynamical interactions with other stars or molecular clouds (\\S~\\ref{surv.sec}). The nearest stars to the Sun, $\\alpha$ Cen and Proxima Cen, form a wide pair which may be on a hyperbolic orbit \\citep{ano}. It is expected that such systems should be mostly young, or belong to moving groups, remnant clusters or associations, but this has not yet been demonstrated on a representative sample. It is not known if the companions formed together and have the same age. We combine age-related parameters and data, including color-absolute magnitude diagrams (\\S~\\ref{hr.sec}), chromospheric activity indeces (\\S~\\ref{chr.sec}), coronal X-ray luminosity (\\S~\\ref{x.sec}), multiplicity parameters (\\S~\\ref{bin.sec}) and kinematics (\\S~\\ref{skg.sec}) to shed light on this problem. Previous investigations in the field are too numerous to be listed, but a few papers in considerable overlap with this study should be mentioned. \\citet{pov} published a catalog of 305 nearby wide binary and 29 multiple systems. They discussed the importance of moving groups for separating different species of wide binaries and tentatively assigned 32 systems to the Hyades stream (called supercluster following Eggen's nomenclature), and 14 to the Sirius stream. \\citet{sal} undertook a comprehensive revision of the Luyten catalog for approximately 44\\% of the sky, drastically improving precision of epoch 2000 positions and proper motions, and supplying the stars with NIR magnitudes from 2MASS. This allowed \\citet{gou} to estimate, for the first time, trigonometric parallaxes of 424 common proper motion companions to Hipparcos stars with reliable parallaxes. This extrapolation of parallaxes to CPM companions is justified for high-proper motion pairs where the rate of chance alignments is small. There is significant overlap between the sample investigated in this paper and the catalog of \\citet{gou}, although we did not use the latter as a starting point for our selection. We are also employing the parallax extrapolation technique for dim companions not observed by Hipparcos when constructing color-magnitude diagrams in this paper. ", "conclusions": "\\label{dis.sec} One of the most interesting results of this paper is that we find little, if any, presence of thick disk or halo population in the local sample of very wide binaries. The only CPM system that may belong to the thick disk is the WD+dM4.5 pair HIP 65877 (DA3.5 white dwarf WD $1327-083$) and LHS 353, cf. \\citet{sil}. This shows that even the widest pairs at separations greater than $1000$ AU can survive for $\\simeq$ 1 Gyr staying constantly in the thin disk of the Galaxy, despite numerous encounter and dynamical interaction events. This observation does not refute the dynamical analysis presented in \\S~\\ref{surv.sec}, because thick disk and halo stars are very rare in the Solar vicinity, and our sample (based on bright stars in the Hipparcos catalog) is probably too small and incomplete to accommodate sufficient statistics. But if we boldly extrapolate this result to a wider part of the Galaxy, we conclude that normal thin disk, moderately young or very young, stars dominate wide CPM binary and multiple systems. Statistically, this is quite consistent with the estimation by \\citet{bart} on a larger sample of 804 Hipparcos visual systems, who found that $92\\%$ of systems belong to the thin disk (and are mostly young to middle-age), $7.6\\%$ to the thick disk, and much less than $1\\%$ to the halo. Further inroads in this study can be made by collecting a larger volume-limited sample of very wide binaries and a comparison with a representative set of nearby field stars. We consider this paper as an initial step in this direction. Despite the considerable progress in recent years, chronology of solar-type stars is still in a rather pitiful state, and we find more evidence to this in the widely discrepant age estimates for a few CPM systems obtained with different methods. Although a significant fraction of CPM companions display enhanced chromospheric, X-ray and EUV activity, only few are patently in the pre-main-sequence stage of evolution (e.g., AT and AU Mic) where these signs of activity and the high rate of surface rotation can be attributed to a very young age. The origin of activity in most of our CPM systems lies in short-period binarity of their components, i.e., in hierarchical multiplicity. An interesting connection emerges between the presence of wide companions and the existence of short-period binaries. The reason for abundant multiple systems may partly be purely dynamical, in that the chances of survival are higher for systems with an internal binary because of the larger mass. An alternative astrophysical possibility is that the original fragmentation of a star-forming core takes place at various spatial scales and tends to produce multiple stellar systems, of which only hierarchical ones can survive for an appreciably long time. Apparently, the time-scale of dynamical survival of wide companions (of order 1 Gyr) is sufficiently long compared to the time-scale of dynamical evolution of non-coplanar multiple systems (the Kozai cycle, \\S \\ref{kozai.sec}) for the latter to shape up the present-day systems. The existence of circularized spectroscopic binaries with periods less than a few days may be the direct consequence of the interaction with remote companions, followed by the tidal friction and loss of angular momentum \\citep{eggl}. Ultimately, the inner components will form a contact binary and then merge. The existence of CPM multiple systems in a wide range of ages and separations will allow us to investigate this process in detail as it unfolds. Indeed, even in our sample of modest size we find examples of inner pairs of intermediate periods and large eccentricities, which are apparently evolving toward the tidally circularized state. The Kozai-type mechanism can affect the dynamical stability and composition of planetary systems. We find two stars in our sample with multiple planets (55 Cnc and GJ 777 A), and both have interesting dynamical properties very much unlike our Solar system." }, "0808/0808.0371_arXiv.txt": { "abstract": "We report high angular resolution observations of the HCN (3-2) line emission in the circumstellar envelope of the O-rich star W~Hya with the Submillimeter Array. The proximity of this star allows us to image its molecular envelope with a spatial resolution of just $\\sim 40$ AU, corresponding to about 10 times the stellar diameter. We resolve the HCN (3-2) emission and find that it is centrally peaked and has a roughly spherically symmetrical distribution. This shows that HCN is formed in the innermost region of the envelope (within $\\sim 10$ stellar radii), which is consistent with predictions from pulsation-driven shock chemistry models, and rules out the scenario in which HCN forms through photochemical reactions in the outer envelope. Our model suggests that the envelope decreases steeply in temperature and increases smoothly in velocity with radius, inconsistent with the standard model for mass-loss driven by radiative pressure on dust grains. We detect a velocity gradient of $\\sim 5$ km~s$^{-1}$ in the NW--SE direction over the central 40 AU. This velocity gradient is reminescent of that seen in OH maser lines, and could be caused by the rotation of the envelope or by a weak bipolar outflow. ", "introduction": "The detection at radio frequencies of various molecules in the circumstellar envelopes of evolved stars has considerably improved our understanding of circumstellar chemistry. This is particularly true for the prototype of carbon stars, IRC+10216, on which much attention has been focused. Several dozens of molecules have been detected in its circumstellar envelope (e.g., \\citealt{cer00}), reflecting its complex and rich chemistry. The inventory of molecules together with high angular resolution images of their spatial distribution have led to the construction of sophisticated chemical models for the circumstellar envelope of IRC+10216 (e.g., \\citealt{gla87,mil94}) that now form the basis of our understanding of the chemistry in C-rich envelopes. The chemistry in O-rich envelopes, on the other hand, was not expected to be as rich because equilibrium chemistry calculations indicate that nearly all the carbon should be locked into CO and nitrogen into N$_2$ (\\citealt{tsu73}). Indeed, before 1985, the inventory of molecules detected around oxygen stars was sparse and did not include carbon-bearing molecules, except of CO. \\cite{deg85} reported the first detection of HCN towards O-rich stars. The list of HCN detections was later extended by \\cite{jew86}, \\cite{lin88}, \\cite{ner89}, \\cite{olo98} and \\cite{bie00}. Recently, \\cite{ziu07} also emphasized the chemical complexity in the envelope of the O-rich supergiant star VY~CMa, following the detection of various molecules such as HCO$^+$, CS, NaCl, PN and SiS. Two competing models have been proposed to account for the formation of HCN in O-rich envelopes: photochemical reactions in the outer envelope (\\citealt{cha95}), or gas-phase non-equilibrium chemical reactions in the inner region close to the stellar photosphere (within few tens of AU from the star) due to shocks driven by stellar pulsation (\\citealt{dua99,dua00}). The two models predict very different HCN spatial distribution: in a hollow shell-like structure of radius 200 to 1000 AU depending on the mass loss for photochemical reactions, or a centrally concentrated distribution extending up to the photodissociation radius of HCN for the shock-driven chemical reactions. Current observational data seem to favor pulsation-driven shock chemistry formation of HCN in O-rich stars. \\cite{bie00} conducted single-dish observations toward a sample of 16 O-rich stars in the HCN (3-2) and (4-3) lines. The detection of these high density tracers is inconsistent with photochemical production of HCN in the outer envelope. The detection of the HCN (3-2) (0,1$^{1c}$,0) and (8-7) transitions toward $\\chi$ Cyg gave further support for the formation of HCN in the innermost region of the envelope, within $\\sim 20$ stellar radii (i.e. $\\sim 30$ AU), based on excitation arguments (\\citealt{dua00}). \\cite{mar05} reported the first interferometric observations of HCN in the circumstellar envelopes of the O-rich stars IK~Tau and TX~Cam. In both stars, the HCN emission appears to be largely concentrated within $\\sim 800$ AU, as predicted by shock chemistry models. If HCN indeed forms close to the stellar photosphere, it should be an excellent tracer of the inner envelope, and perhaps even of the wind acceleration zone where, in the standard model, dust particles form and radiation pressure on dust becomes most effective. Maser (e.g., H$_2$O or OH) emission in the inner envelope can be observed with very long baseline interferometry, but traces only a limited radial range. On the other hand, there are so far only few reports of high angular resolution observations of other thermal lines in the inner envelope, such as SiO (2-1) $\\nu = 0$ (\\citealt{luc92,sah93,sch04}), none of them with angular resolution higher than $\\sim 100$ AU. In this letter, we report very high spatial resolution ($\\sim 40$ AU) observations of the HCN (3-2) emission toward W~Hya with the Submillimeter Array. W~Hya is one of the closest O-rich stars at a distance of 78 pc (based on Hipparcos data, \\citealt{per97}, revised by \\citealt{kna03}), and shows the brightest HCN (3-2) and (4-3) emission amongst the sample of O-rich stars observed by \\cite{bie00}. It is therefore an excellent target in which to investigate the distribution and kinematics of HCN in the circumstellar envelope of an O-rich star. ", "conclusions": "" }, "0808/0808.0692_arXiv.txt": { "abstract": "We present {\\it Chandra} ACIS X-ray observations of the Galactic supernova remnant Cassiopeia A taken in December 2007. Combining these data with previous archival {\\it Chandra} observations taken in 2000, 2002, and 2004, we estimate the remnant's forward shock velocity at various points around the outermost shell to range between 4200 and 5200 $\\pm500 $ km s$^{-1}$. Using these results together with previous analyses of Cas A's X-ray emission, we present a model for the evolution of Cas A and find that it's expansion is well fit by a $\\rho_{ej} \\propto r^{-(7-9)}$ ejecta profile running into a circumstellar wind. We further find that while the position of the reverse shock in this model is consistent with that measured in the X-rays, in order to match the forward shock velocity and radius we had to assume that $\\sim$ 30\\% of the explosion energy has gone into accelerating cosmic rays at the forward shock. The new X-ray images also show that brightness variations can occur for some forward shock filaments like that seen for several nonthermal filaments seen projected in the interior of the remnant. Spectral fits to exterior forward shock filaments and interior nonthermal filaments show that they exhibit similar spectra. This together with similar flux variations suggests that interior nonthermal filaments might be simply forward shock filaments seen in projection and not located at the reverse shock as has been recently proposed. ", "introduction": "Cassiopeia A (Cas A) is one of the youngest known Galactic supernova remnants (SNR) with an estimated explosion date no earlier than $1681 \\pm19$ \\citep{fesen06}. Optical echoes of the supernova outburst have been recently detected \\citep{Rest08}, the spectra of which indicate Cas~A is the remnant of a Type IIb supernova event \\citep{Krause08} probably from a red supergiant in the mass range of 15--25 M$_{\\odot}$ that may have lost much its hydrogen envelope to a binary interaction \\citep{Young06}. Viewed in X-rays, the remnant consists of a line emitting shell arising from reverse shocked ejecta rich in O, Si, Ar, Ca, and Fe \\citep{fabian80,markert83,vink96,hughes00,will02,will03,hwang03,laming03}. Exterior to this shell are faint X-ray filaments which mark the current position of the remnant's forward shock front. The emission found here is nonthermal X-ray synchrotron radiation as well as faint line emission from shocked circumstellar material (CSM). \\citet{vink98} compared {\\sl Einstein} HRI to {\\sl ROSAT} HRI observations of Cas A to measure the expansion of the bright shell, finding an expansion age of $\\sim$ 500 yr, considerably less than the $\\sim$ 800 yr expansion age derived from 1.5 and 5.0 GHz radio observations \\citep{anderson95}, but similar to the 400--500 yr expansion age found by \\citet{agueros99} using data taken at 151 MHz. More recently, \\citet{delaney03} using {\\sl Chandra} X-ray observations taken in 2000 and 2002 presented the first proper motion measurements of the forward blastwave velocity. Assuming a distance of 3.4 kpc \\citep{reed95}, they estimated a blast wave expansion velocity of $\\approx$ 5000 km s$^{-1}$. Besides the outlying nonthermal emission filaments associated with the forward shock, some filamentary nonthermal X-ray emission is also seen in projection in the interior of the SNR \\citep{delaney04a}. Whether these interior filamentary emissions originate from a wrinkled forward shock seen in projection or arises from nonthermal emission mechanisms in the interior of the SNR is currently uncertain \\citep{laming01,uchiyama08,helder08}. Comparisons of {\\it Chandra} observations taken in 2000, 2002, and 2004 revealed secular changes in several X-ray thermal knots and in one nonthermal filament projected in the remnant's interior \\citep{patnaude07}. \\citet{uchiyama08} using the same multi-epoch {\\it Chandra} observations found evidence for rapid variability in many more interior nonthermal X-ray emission filaments. Motivated by similar changes seen in RX J1713-3946 \\citep{uchiyama07}, they measured the time variability of selected filaments to determine the local magnetic field strength in the variable regions. Their results suggest that the magnetic field in these regions is relatively high, B $\\sim$ 1 mG. Such a high magnetic field strength would be consistent with equipartition field strengths inferred in observations of bright radio knots in the remnant \\citep{longair94,wright99}. \\citet{uchiyama08} argue that their result points to a synchrotron origin for the emission coming from these knots, ruling out nonthermal bremsstrahlung from $\\sim$ 100 keV electrons \\citep{laming01}, and suggest that this is strong evidence for a hadronic origin to the TeV emission observed in Cas A \\citep{aharonian01,albert07}. Based on the location of the synchrotron knots, Uchiyama \\& Aharionian suggest that the emission is located primarily at the reverse shock, and \\citet{helder08} reach a similar conclusion. Here we present forward shock velocity measurements using new {\\it Chandra} ACIS observations of Cas A taken in December 2007 and compare these results to models for SNR evolution with and without efficient shock acceleration. The new observations show that many nonthermal emission filaments and features have undergone substantial brightness variations over the last four years. Model fits to the nonthermal emission coming from both the forward shock and the interior filaments indicate that they are quantitatively similar. We also present evidence for fast variability in forward shock front filaments which argues against the conclusion that rapid variability is a property restricted to emission at the reverse shock. ", "conclusions": "We have presented new {\\it Chandra} ACIS observations of Cas A which were taken in late 2007. These new observations, when combined with previous {\\it Chandra} data, allow us to constrain the velocity of the forward shock to be about 4900 km s$^{-1}$. Combined with results from previous analyses of Cas A's X-ray emission \\citep{laming03,gotthelf01}, we present several models for the evolution of Cas A and find that it's expansion can be well modeled by an $n= 7 - 9$ ejecta profile running into a circumstellar wind. We also find that the position of the reverse shock in this model is consistent with that measured by \\citet{gotthelf01}. However, in order to match the radius of the forward shock, we found that we must assume that the forward shock is efficiently accelerating cosmic rays. Rapid changes in Cas A's synchrotron emission are seen for interior and exterior projected filaments, with both showing similar nonthermal spectra as well as inferred magnetic field strengths. Based on this and the simulations presented by \\citet{bykov08}, it is currently not clear whether the interior filaments are in fact located at the reverse shock as recently argued by \\citet{uchiyama08} and \\citet{helder08}. Instead, we propose that the interior filaments might be forward shocks seen in projection \\citep{delaney04b}. In that case, the observed brightness variations might arise from wrinkles in front-facing, forward shock as it moves through an inhomogeneous, local circumstellar medium. Although we cannot rule out the possibility that interior nonthermal filaments are associated with the reverse shock, the combination of similar spectra, flaring timescale, and our fits to the remnant's dynamics are suggestive that the observed synchrotron flaring for interior filaments arises from forward shock filaments seen in projection toward Cas A's interior rather than at the reverse shock as recently suggested. At the least, our new X-ray data of Cas A shows that rapid brightness variations like those seen for interior nonthermal filaments can also be exhibited by some outer, nonthermal forward shock filaments." }, "0808/0808.2973_arXiv.txt": { "abstract": "We present high-angular resolution sub-millimeter continuum images and molecular line spectra obtained with the Submillimeter Array toward two massive cores that lie within Infrared Dark Clouds; one actively star-forming (G034.43+00.24~MM1) and the other more quiescent (G028.53$-$00.25~MM1). The high-angular resolution sub-millimeter continuum image of G034.43+00.24~MM1 reveals a compact ($\\sim$ 0.03 pc) and massive ($\\sim$ 29\\,\\Msun) structure while the molecular line spectrum shows emission from numerous complex molecules. Such a rich molecular line spectrum from a compact region indicates that G034.43+00.24~MM1 contains a hot molecular core, an early stage in the formation of a high-mass protostar. Moreover, the velocity structure of its $^{13}${\\rmfamily CO}{\\,(3--2)} emission indicates that this B0 protostar may be surrounded by a rotating circumstellar envelope. In contrast, the sub-millimeter continuum image of G028.53$-$00.25~MM1 reveals three compact ($\\lesssim$ 0.06 pc), massive (9--21\\,\\Msun) condensations but with no lines detected in its spectrum. We suggest that the core G028.53$-$00.25~MM1 is in a very early stage in the high-mass star-formation process; its size and mass are sufficient to form at least one high-mass star, yet it shows no signs of localized heating. Because the combination of high velocity line wings with a large IR--mm bolometric luminosity ($\\sim$~10$^{2}$\\,\\Lsun) indicates that this core has already begun to form accreting protostars, we speculate that the condensations may be in the early phase of accretion and may eventually become high-mass protostars. We, therefore, have found the possible existence of two high-mass star-forming cores; one in a very early phase of star-formation and one in the later hot core phase. Together the properties of these two cores support the idea that the earliest stages of high-mass star-formation occur within IRDCs. ", "introduction": "Emerging evidence suggests that the very earliest stages of high-mass star and cluster formation occur within cold, dense molecular clumps called infrared dark clouds (IRDCs; \\citealp{Simon-msxgrs,Rathborne06,Rathborne07}). Because these molecular clumps have very high column densities ($\\sim 10^{23}$--$10^{25}$ cm$^{-2}$) and low temperatures ($<25$ K), they have predominantly been identified via their absorption of the background Galactic emission at mid-IR wavelengths (e.g., \\citealp{Egan98,Carey98,Carey00,Simon-catalog}). Because IRDCs are the densest clumps in molecular clouds \\citep{Simon-msxgrs} which are now undergoing the process of fragmentation and condensation, IRDCs are important laboratories to study the pristine, undisturbed physical conditions of cluster-forming clouds before they are shredded apart by stellar winds and radiation. Millimeter/sub-millimeter continuum studies of IRDCs \\citep{Lis94,Carey00,Garay04,Ormel05,Beuther05,Rathborne05,Rathborne06} reveal that IRDCs contain many compact cores. These cores have typical sizes of $<$~0.5~pc and masses of $\\sim$~120\\,\\Msun\\, \\citep{Rathborne06}. While most of these cores contain little evidence for active star formation, some are associated with bright 24\\,\\um\\, emission, broad molecular line emission, shocked gas, and maser emission, indicating that they are actively forming stars (e.g., \\citealp{Rathborne05,Wang06,Chambers08}). Indeed, a number of low- and intermediate-mass protostars \\citep{Carey00,Redman03} in addition to high-mass protostars \\citep{Beuther05,Rathborne05,Pillai06,Wang06} have been identified within IRDCs. Hot Molecular Cores (HMCs) are associated with the early stages of high-mass star formation. They correspond to the stage immediately after a cold dense starless core has formed. During the later stages in their evolution, HMCs are often also associated with methanol masers and ultra-compact \\hii\\, regions. Numerous examples of HMCs have been found throughout the Galaxy (e.g., \\citealp{Garay99,Kurtz00,Churchwell02}), including one within an IRDC \\citep{Rathborne07}. HMCs are compact ($<$0.1 pc), dense ($\\sim$10$^{5}-10^{8}$\\,\\cmc), and massive ($\\sim$10$^{2}$\\,\\Msun;\\citealp{Garay99,Kurtz00,Churchwell02}). Due to the internal heating from the central high-mass protostar (T$\\sim$100\\,K), they are typically very luminous ($>$10$^3$\\,\\Lsun) and show strong emission from complex molecules (e.g., \\citealp{Kurtz00}). Accretion disks are also inferred in the later stage of the HMC phase directly \\citep{Zhang05a,Cesaroni07} or indirectly \\citep{Kurtz00,Beuther02,Zhang01,Zhang05b} by the presence of molecular outflows and maser emission. The existence of HMCs in IRDCs establishes a possible link between IRDCs and the early stages of high-mass star formation \\citep{Rathborne07}. To understand the earliest stages in high-mass star formation one needs to identify and study the so-called `high-mass starless cores,' the immediate precursors to high-mass protostars, HMCs, and ultra-compact \\hii\\, regions. High-mass star formation is rare and occurs rapidly; thus, the identification of high-mass starless cores is difficult. Because IRDCs have sizes, masses, and densities similar to cluster-forming molecular clumps but are considerably colder, we suggest that they are the precursors to clusters and their dense, cold cores the precursors to stars. Thus, the high-mass cores within IRDCs that contain no evidence for star formation are good candidates for the elusive `high-mass starless cores.' Here we present interferometric sub-millimeter continuum and molecular line observations obtained with the Submillimeter Array toward two cores within IRDCs; one showing evidence for active high-mass star formation (\\irdcfortythree), the other showing none (\\irdcthirty). Both cores are at distances d $>$ 3.7\\,kpc and have sizes R $<$0.8 pc, gas masses M $>$400\\,\\Msun, and bolometric luminosities L $>$10$^{2}$\\,\\Lsun\\, (Table~\\ref{core-properties}; \\citealp{Simon-msxgrs,Rathborne06,Rathborne08}). We speculate that the `active' core is in a later evolutionary stage than the more quiescent core. The new interferometric data reveal distinct properties for these cores and suggest that \\irdcfortythree\\, is a more evolved high-mass protostar in the HMC phase and that \\irdcthirty\\, may be in a very early `starless' core phase of high-mass star-formation. ", "conclusions": "Using the SMA we have obtained high-angular resolution sub-millimeter continuum images and molecular line spectra toward two cores within IRDCs. In the low-angular resolution data these cores have sizes R $<$0.8\\,pc, gas masses M $>$400\\,\\Msun, and bolometric luminosities L~$\\gtrsim$10$^{2}$\\,\\Lsun\\, \\citep{Rathborne06,Rathborne08}. Despite their similar sizes, masses, and bolometric luminosities the two cores show remarkable differences in the high-angular resolution data. The sub-millimeter continuum images reveal that \\irdcfortythree\\,remains unresolved and has a size of $\\sim$ 0.03 pc and mass of 29\\,\\Msun. In contrast, \\irdcthirty\\, contains at least three compact ($\\sim$ 0.06 pc), massive (9--21\\,\\Msun) condensations. Moreover, the molecular line spectra reveal that \\irdcfortythree\\, emits strong lines from many complex molecules, but the core \\irdcthirty\\, does not. This suggests that \\irdcfortythree\\, is hot, compact, and dense, while \\irdcthirty\\, is colder and possibly more extended. We speculate, therefore, that \\irdcfortythree\\, is a HMC in an early stage in the formation of a high-mass protostar, while \\irdcthirty\\, may represent an even earlier phase in the high-mass star-formation process. Because \\irdcthirty\\, shows some evidence for active star formation in the low-angular resolution data (e.g., shocked gas, high velocity line profiles, and a high bolometric luminosity), we cannot rule out the possibility that the condensations within this core may have already begun to accrete. However, the low temperatures and absence of bright 24\\,\\um\\, point sources or strong molecular line emission from high excitation transitions suggests that there is little localized heating. Because of its size and mass, \\irdcthirty\\, is likely to give rise to at least one high-mass star. We speculate that the detected condensations may continue to accrete material from their surroundings to eventually form high-mass stars. Since the condensations account for only $\\sim$4\\% of the total mass, there is an ample amount of supply of material to accrete. The properties of these cores support the idea that high-mass star and cluster formation likely occurs in IRDCs. High-angular sub-millimeter continuum images and molecular line spectra toward more cores within IRDCs are needed to unambiguously establish the evolutionary path from a high-mass starless core to a high-mass protostar." }, "0808/0808.2718_arXiv.txt": { "abstract": "We study the effects of Supernova (SN) feedback on the formation of galaxies using hydrodynamical simulations in a $\\Lambda$CDM cosmology. We use an extended version of the code GADGET-2 which includes chemical enrichment and energy feedback by Type II and Type Ia SN, metal-dependent cooling and a multiphase model for the gas component. We focus on the effects of SN feedback on the star formation process, galaxy morphology, evolution of the specific angular momentum and chemical properties. We find that SN feedback plays a fundamental role in galaxy evolution, producing a self-regulated cycle for star formation, preventing the early consumption of gas and allowing disks to form at late times. The SN feedback model is able to reproduce the expected dependence on virial mass, with less massive systems being more strongly affected. ", "introduction": "Supernova explosions play a fundamental role in galaxy formation and evolution. On one side, they are the main source of heavy elements in the Universe and the presence of such elements substantially enhances the cooling of gas (White \\& Frenk 1991). On the other hand, SNe eject a significant amount of energy into the interstellar medium. It is believed that SN explosions are responsible of generating a self-regulated cycle for star formation through the heating and disruption of cold gas clouds, as well as of triggering important galactic winds such as those observed (e.g. Martin 2004). Smaller systems are more strongly affected by SN feedback, because their shallower potential wells are less efficient in retaining baryons (e.g. White \\& Frenk 1991). Numerical simulations have become an important tool to study galaxy formation, since they can track the joint evolution of dark matter and baryons in the context of a cosmological model. However, this has shown to be an extremely complex task, because of the need to cover a large dynamical range and describe, at the same time, large-scale processes such as tidal interactions and mergers and small-scale processes related to stellar evolution. One of the main problems that galaxy formation simulations have repeteadly found is the inability to reproduce the morphologies of disk galaxies observed in the Universe. This is generally refered to as the angular momentum problem that arises when baryons transfer most of their angular momentum to the dark matter components during interactions and mergers (Navarro \\& Benz 1991; Navarro \\& White 1994). As a result, disks are too small and concentrated with respect to real spirals. More recent simulations which include prescriptions for SN feedback have been able to produce more realistic disks (e.g. Abadi et al. 2003; Robertson et al. 2004; Governato et al. 2007). These works have pointed out the importance of SN feedback as a key process to prevent the loss of angular momentum, regulate the star formation activity and produce extended, young disk-like components. In this work, we investigate the effects of SN feedback on the formation of galaxies, focusing on the formation of disks. For this purpose, we have run simulations of a Milky-Way type galaxy using an extended version of the code {\\small GADGET-2} which includes chemical enrichment and energy feedback by SN. A summary of the simulation code and the initial conditions is given in Section~\\ref{simus}. In Section~\\ref{results} we investigate the effects of SN feedback on galaxy morphology, star formation rates, evolution of specific angular momentum and chemical properties. We also investigate the dependence of the results on virial mass. Finally, in Section~\\ref{conclusions} we give our conclusions. ", "conclusions": "We have run simulations of a Milky Way-type galaxy in its cosmological setting in order to investigate the effects of SN feedback on the formation of galaxy disks. We compare two simulations with the only difference being the inclusion of the SN energy feedback model of Scannapieco et al. (2005, 2006). Our main results can be summarized as follows: \\begin{itemize} \\item{ SN feedback helps to settle a self-regulated cycle for star formation in galaxies, through the heating and disruption of cold gas and the generation of galactic winds. The regulation of star formation allows gas to be mantained in a hot halo which can condensate at late times, becoming a reservoir for recent star formation. This contributes significantly to the formation of disk components. } \\item{When SN feedback is included, the specific angular momentum of the baryons is conserved and disks with the correct scale-lengths are obtained. This results from the late collapse of gas with high angular momentum, which becomes available to form stars at later times, when the system does not suffer from strong interactions. } \\item{ The injection of SN energy into the interstellar medium generates a redistribution of chemical elements in galaxies. If energy feedback is not considered, only the very central regions were stars are formed are contaminated. On the contrary, the inclusion of feedback triggers a redistribution of metals since gas is heated and expands, contaminating the outer regions of galaxies. In this case, metallicity profiles in agreement with observations are produced. } \\item{ Our model is able to reproduce the expected dependence of SN feedback on virial mass: as we go to less massive systems, SN feedback has stronger effects: the star formation rates (normalized to mass) are lower, and more violent winds develop. This proves that our model is well suited for studying the cosmological growth of structure where large systems are assembled through mergers of smaller substructures and systems form simultaneously over a wide range of scales. } \\end{itemize}" }, "0808/0808.0001_arXiv.txt": { "abstract": "We present photometry for globular and open cluster stars observed with the Sloan Digital Sky Survey (SDSS). In order to exploit over 100 million stellar objects with $r < 22.5$~mag observed by SDSS, we need to understand the characteristics of stars in the SDSS $ugriz$ filters. While star clusters provide important calibration samples for stellar colors, the regions close to globular clusters, where the fraction of field stars is smallest, are too crowded for the standard SDSS photometric pipeline to process. To complement the SDSS imaging survey, we reduce the SDSS imaging data for crowded cluster fields using the DAOPHOT/ALLFRAME suite of programs and present photometry for 17 globular clusters and 3 open clusters in a SDSS value-added catalog. Our photometry and cluster fiducial sequences are on the native SDSS 2.5-meter $ugriz$ photometric system, and the fiducial sequences can be directly applied to the SDSS photometry without relying upon any transformations. Model photometry for red giant branch and main-sequence stars obtained by Girardi et~al.\\ cannot be matched simultaneously to fiducial sequences; their colors differ by $\\sim0.02$--$0.05$~mag. Good agreement ($\\la0.02$~mag in colors) is found with Clem et~al.\\ empirical fiducial sequences in $u'g'r'i'z'$ when using the transformation equations in Tucker et~al. ", "introduction": "As single-age and (in most cases) single-metallicity populations, Galactic star clusters provide important calibration samples for exploring the relationships between stellar colors and absolute magnitudes as functions of stellar age and heavy-element content. These two observable properties of a star are related to fundamental physical parameters, such as the effective temperature ($T_{\\rm eff}$) and surface gravity ($\\log{g}$), as well as the metallicity. The color and magnitude relations can be used to test stellar evolutionary theories, to interpret the observed distribution of stars in color-color and color-magnitude diagrams (CMDs), and to derive distances to stars and star clusters via photometric parallax or main-sequence (MS) fitting techniques \\citep[e.g.,][]{johnson:57}. Because the relationships between magnitude, color, and fundamental stellar properties depend on the filters used, it is necessary to characterize these relations for each filter system. Galactic globular and open clusters provide an ideal opportunity to achieve this goal because the same distance can be assumed for cluster members with a wide range of stellar masses. Furthermore, observations of a large number of Galactic clusters can cover a wide range of the heavy-element content, providing an opportunity to explore the effects of metallicity on magnitudes and colors for each set of filter bandpasses. Among previous and ongoing optical surveys, the Sloan Digital Sky Survey \\citep[SDSS;][]{york:00,edr,dr1,dr2,dr3,dr4,dr5,dr6} is the largest and most homogeneous database of stellar brightnesses currently available. The original goal of the SDSS was to survey large numbers of galaxies and quasars. However, in the first five years of operation, SDSS-I has made remarkable contributions to our understanding of the Milky Way and its stellar populations \\citep[e.g.,][]{newberg:02,allendeprieto:06,belokurov:06,dejong:08,juric:08}. These successes have initiated the Galactic structure program SEGUE (Sloan Extension for Galactic Understanding and Exploration; B.\\ Yanny et al.\\ 2008, in preparation), one of the surveys being conducted in the ongoing three year extension of the survey (SDSS-II). When SDSS-II finishes, it will provide imaging data for approximately 10,000 square degrees of the northern sky. SDSS measures the brightnesses of stars using a dedicated 2.5-m telescope \\citep{gunn:06} in five broadband filters $u$, $g$, $r$, $i$, and $z$, with average wavelengths of 3551\\AA, 4686\\AA, 6165\\AA, 7481\\AA, and 8931\\AA, respectively \\citep{fukugita:96,edr}. The 95\\% detection repeatability limits are 22.0~mag, 22.2~mag, 22.2~mag, 21.3~mag, and 20.5~mag for point sources in $u$, $g$, $r$, $i$, and $z$, respectively. The rms photometric precision is $0.02$~mag for sources not limited by photon statistics \\citep{ivezic:03}, and the photometric calibration is accurate to $\\sim2\\%$ in the $g$, $r$, $i$ bands, and $\\sim3\\%$ in $u$ and $z$ \\citep{ivezic:04}. The SDSS filters represent a new filter set for stellar observations, and therefore it is important to understand the properties of stars in this system. Furthermore, future imaging surveys such as the Panoramic Survey Telescope \\& Rapid Response System \\citep[Pan-STARRS;][]{kaiser:02} and the Large Synoptic Survey Telescope \\citep[LSST;][]{stubbs:04} will use similar photometric bandpasses, providing even deeper data in $ugriz$ than SDSS over a larger fraction of the sky. During the course of SDSS-I, about 15 globular clusters and several open clusters were observed. Several more clusters were imaged in SDSS-II including M71. These clusters together provide accurate calibration samples for stellar colors and magnitudes in the SDSS filters. The SDSS images are processed using the standard SDSS photometric pipelines \\citep[{\\it Photo};][]{lupton:02}. {\\it Photo} pre-processes the raw images, determines the point spread function (PSF), detects objects, and measures their properties. Photometric calibration is then carried out using observations of stars in the secondary patch transfer fields \\citep{tucker:06,davenport:07}. In this paper, we simply refer to these calibrated magnitudes as {\\it Photo} magnitudes. \\begin{figure*} \\epsscale{0.9} \\plotone{f1.ps} \\caption{CMDs of M3 from the SDSS photometric pipeline ({\\it Photo}; {\\it left}) and DAOPHOT reduction in this paper ({\\it right}). Stars within a $30\\arcmin$ radius from the cluster center are shown in the left panel, but the {\\it Photo} photometry is only available on the outskirts of the cluster. In the right panel, RR Lyraes are scattered off the cluster horizontal branch. \\label{fig:photo}} \\end{figure*} {\\it Photo} was originally designed to handle high Galactic latitude fields with relatively low densities of Galactic field stars (owing to the primarily extragalactic mission of SDSS-I); however, there are some concerns about its photometry derived in crowded fields \\citep{dr6}. In particular, stellar clusters present a challenge to {\\it Photo}. Firstly, {\\it Photo} slows down dramatically in the high density cluster cores, which are too crowded for {\\it Photo} to process, so it does not provide photometry for the most crowded regions of these scans. Figure~\\ref{fig:photo} compares a CMD for the globular cluster M3 obtained from {\\it Photo} photometry to that obtained from a DAOPHOT \\citep{stetson:87} reduction in this paper, which is specifically designed for crowded field photometry. The {\\it Photo} photometry is only available on the outskirts of the cluster, and it provides a considerably less well-defined subgiant branch (SGB), red giant branch (RGB), and horizontal branch (HB). Secondly, there is a concern that the photometry in the area surrounding clusters, and in low Galactic latitude fields, may also be affected by inaccurate modeling of the PSF if stars in crowded regions were selected as PSF stars by {\\it Photo}. Photometric information in crowded fields can be extracted from the original SDSS imaging data. For example, \\citet{smolcic:07} used the DoPHOT \\citep{schechter:93} photometry package to explore the structure of the Leo~I dwarf spheroidal galaxy (dSph). Similarly, \\citet{coleman:07} used the DAOPHOT package to study the stellar distribution of the dSph Leo~II. In this paper, we employ the DAOPHOT/ALLFRAME \\citep{stetson:87,stetson:94} suite of programs to derive photometry for 17 globular clusters and 3 open clusters that have been observed with SDSS. We derive photometry by running DAOPHOT for SDSS imaging frames where {\\it Photo} did not run. In addition, we reduce imaging data for fields farther away from the clusters, where the {\\it Photo} results are expected to be reliable, in order to set up photometric zero points for the DAOPHOT photometry. We also compare DAOPHOT and {\\it Photo} results for the open cluster fields to verify the accuracy of the {\\it Photo} magnitudes in these semi-crowded fields. An overview of the SDSS imaging survey and our sample clusters are presented in \\S~2. In \\S~3 we describe the preparation of imaging data from the SDSS database. In \\S~4 we describe the method of crowded field photometry using DAOPHOT/ALLFRAME, and evaluate the photometric accuracy. In \\S~5 we present cluster photometry and fiducial sequences, and compare them with theoretical stellar isochrones and fiducial sequences in $u'g'r'i'z'$. ", "conclusions": "" }, "0808/0808.3366_arXiv.txt": { "abstract": "The relevance of storage-ring electron-ion recombination experiments for astro\\-physics is outlined. In particular, the role of low-energy dielectronic-recombination resonances is discussed. A bibliographic compilation of electron-ion recombination measurements with cosmically abundant ions is provided. ", "introduction": "Heavy-ion storage rings equipped with electron coolers are an excellent experimental environment for electron-ion collision studies. Some recent studies of dielectronic recombination (DR) focussed on high-resolution spectroscopy of highly-charged ions. Highlights of this research are the measurement of the hyperfine induced decay rate of the $1s^2\\,2s\\,2p\\;^3P_0$ state in berylliumlike Ti$^{18+}$ \\cite{Schippers2007a} utilizing DR at the storage ring TSR of the Heidelberg Max-Planck-Institute for Nuclear Physics, the observation of the isotope shift in DR of three-electron Nd$^{57+}$ \\cite{Brandau2008a} using different isotopes of this ion at GSI's storage ring ESR and the observation of the hyperfine splitting of Sc$^{18+}$ low-energy DR resonances at the TSR high-resolution electron target \\cite{Wolf2006c}. The latter experiment resulted in the determination of the Sc$^{18+}$($2s_{1/2} - 2p_{3/2}$) energy splitting with an uncertainty of only 4.6 ppm which is less than 1\\% of the few-body effects on radiative corrections \\cite{Lestinsky2008a}. Since these exciting developments have already been reviewed recently \\cite{Schippers2008a}, the present review focusses on the relevance of storage-ring electron-ion experiments for astrophysics. Storage-ring experiments provide particularly valuable information on DR in low-temperature plasmas such as photoionized plasmas that occur, e.g., in active galactic nuclei (AGN) in the vicinity of super-massive black holes \\cite{Savin2007d}. In such plasmas highly charged ions exist at relatively low temperatures. For many ions, the DR rate coefficients, that determine the charge balance in these plasmas, depend sensitively on the low-energy DR resonance structure at relative electron-ion energies below $\\lesssim$ 3 eV. In the following, the influence of low-energy DR resonances on electron-ion recombination rate coefficients in low-density plasmas and recent efforts of building a recombination data base for astrophysical modeling of photoionized plasmas are briefly discussed. Finally, a compilation of experimental results for the astrophysically most abundant ions is presented. ", "conclusions": "" }, "0808/0808.3150_arXiv.txt": { "abstract": "We show that the mass-segregation solution for the steady state distribution of stars around a massive black hole (MBH) has two branches: the known weak segregation solution \\citep{bah+77}, and a newly discovered strong segregation solution, presented here. The nature of the solution depends on the heavy-to-light stellar mass ratio $M_{H}/M_{L}$ and on the unbound population number ratio $N_{H}/N_{L}$, through the relaxational coupling parameter $\\Delta\\!=\\!4N_{H}M_{H}^{2}\\left/\\left[N_{L}M_{L}^{2}(3\\!+\\! M_{H}/M_{L})\\right]\\right.$. When the heavy stars are relatively common ($\\Delta\\!\\gg\\!1$), they scatter frequently on each other. This efficient self-coupling leads to weak mass segregation, where the stars form $n\\!\\propto\\! r^{-\\alpha_{M}}$ mass-dependent cusps near the MBH, with indices $\\alpha_{H}\\!=\\!7/4$ for the heavy stars and $3/2\\!<\\!\\alpha_{L}\\!<\\!7/4$ for the light stars (i.e. $\\max(\\alpha_{H}\\!-\\!\\alpha_{L})\\!\\simeq\\!1/4$). However, when the heavy stars are relatively rare ($\\Delta\\!\\ll\\!1$), they scatter mostly on light stars, sink to the center by dynamical friction and settle into a much steeper cusp with $2\\!\\lesssim\\!\\alpha_{H}\\!<\\!11/4$, while the light stars form a $3/2\\!<\\!\\alpha_{L}\\!<\\!7/4$ cusp, resulting in strong segregation (i.e. $\\max(\\alpha_{H}\\!-\\!\\alpha_{L})\\!\\simeq\\!1$). We show that the present-day mass function of evolved stellar populations (coeval or continuously star forming) with a universal initial mass function, separate into two distinct mass scales, $\\sim\\!1\\,\\Mo$ of main sequence and compact dwarfs, and $\\sim\\!10\\,\\Mo$ of stellar black holes (SBHs), and have $\\Delta\\!<\\!0.1$. We conclude that it is likely that many relaxed galactic nuclei are strongly segregated. We review indications of strong segregation in observations of the Galactic Center and in results of numeric simulations, and briefly list some possible implications of a very high central concentration of SBHs around a MBH. ", "introduction": "\\label{s:intro} Early theoretical studies of the dynamics and distribution of stars around a MBH \\citep{pee72,bah+76,bah+77,you80} were triggered by the discovery of quasars \\citep[e.g.][]{mat+63,sch63} and the realization that many galactic nuclei may contain a central massive collapsed object \\citep{lyn69,wol+70}, % {} as well as by the discovery of X-ray sources in globular clusters \\citep{gia+72}, which were then thought to be accreting MBHs (e.g. \\citealt{wyl70,bah+76}; see also \\citealt{mil+02b} for a recent reevaluation of this possibility). The main motivations for these studies were the prospect of detecting MBHs by the observed stellar density profile and by tidal disruption flares, and the possible role of tidal disruptions of stars in the growth of MBHs \\citep{lig+77,coh+78,sha+78,ree88} The renewed interest in this problem is driven by observations of stars closely orbiting the Galactic MBH \\citep{eis+05,ghe+05} and the accumulating data on their distribution and dynamics \\citep{ale99,gen+00,gen+03a,sch+03,sch+07}, as well as by the prospects of detecting gravitational waves (GW) from extreme mass ratio inspiral sources by future GW detectors (EMRIs: compact remnants inspiraling into MBHs, see review by \\citealt{ama+07}; \\citealt{hop06}). EMRI rates and properties strongly depend on the stellar density and the stellar dynamical processes within $O(0.01\\,\\mathrm{pc})$ of the MBH, where inspiraling sources originate \\citep[e.g.][]{hop+05,hop+06a}. Mass segregation occurs in dynamically relaxed systems. MBHs are naturally expected to lie in relaxed cores in scenarios where the MBH is formed by run-away mergers in the extreme central density following core collapse \\citep{ree84}, which occurs on timescales much longer then the relaxation time, $T_{R}$\\texttt{\\textbf{ }}\\citep[e.g. ][]{spi87,qui96b,fre+06b,fre+06c}. Likewise, the extreme mass ratio targets of the planned Laser Interferometer Space Antenna% \\footnote{See LISA mission website\\texttt{ http://lisa.nasa.gov}% } GW detector (LISA) are expected to originate in relaxed nuclei, since LISA design is sensitive to GW from MBHs with mass $\\lesssim\\!10^{7}\\, M_{\\odot}$. The observed correlation between the MBH mass $\\Mbh$ and the typical velocity dispersion of the spheroid of the host galaxy, $\\Mbh\\!\\propto\\!\\sigma^{\\beta}$, $4\\!\\lesssim\\!\\beta\\!\\lesssim5$ \\citep[the $\\Mbh/\\sigma$ relation, ][]{fer+00,geb+00} then implies that such nuclei are dynamically relaxed and very dense \\citep{ale07,mer+07}. This can be seen by assuming for simplicity $\\beta\\!=\\!4$ (a higher value only reinforce these conclusions), and noting that the MBH radius of influence $r_{h}\\!\\sim\\! G\\Mbh/\\sigma^{2}\\!\\propto\\!\\!\\Mbh^{1/2}$ encompasses a stellar mass of order $\\Mbh$, so that the number of stars there is $N_{h}\\!\\sim\\! M_{\\bullet}/\\Ms$, where $\\Ms$ is the typical stellar mass, and the mean stellar density is $\\bar{n}_{h}\\!\\sim\\! N_{h}/r_{h}^{3}\\!\\propto\\! M_{\\bullet}^{-1/2}$. The {}``$nv\\Sigma$'' rate estimate of strong gravitational collisions then implies that $T_{R}^{-1}(r_{h})\\!\\sim\\!\\bar{n}_{h}\\sigma(G\\Ms/\\sigma^{2})^{2}\\!\\propto\\!\\Mbh^{-5/4}$. A more rigorous estimate shows that for the Galactic MBH ($\\Mbh\\!\\simeq\\!4\\times10^{6}\\,\\Mo$, \\citealt{eis+05,ghe+05}), an archetype of LISA targets, $T_{R}\\!\\sim\\! O(1\\,\\mathrm{Gyr})\\!}M_{L}/M_{H}$. However, these two instabilities are distinct effects. The Spitzer instability occurs only when the heavy stars are relatively common in the population, $N_{H}M_{H}^{5/2}/N_{L}M_{L}^{5/2}>\\beta$, where $\\beta\\!\\simeq0.16$ (for $M_{H}\\!\\gg\\! M_{L}$ and $N_{H}M_{H}\\!\\ll\\! N_{L}M_{L}$). In contrast, in the Keplerian potential near a MBH, the Jeans equation dictates that $\\sigma_{H}^{2}/\\sigma_{L}^{2}\\T{=}(5/2+p_{L})/(5/2+p_{H})\\T{\\sim}1$ \\citep{ale+01a}, and so equipartition is never achieved, irrespective of the heavy-to-light mass or number ratios. Strong mass segregation is an instability in the spatial distribution of the heavy stars, which occurs in the opposite limit to the Spitzer instability, when the heavy stars are relatively rare in the population, $\\Delta\\T{\\sim}N_{H}M_{H}/N_{L}M_{H}\\T{<}1$ (for $M_{H}\\T{\\gg}M_{L}$, Eq. \\ref{e:delta}). \\citet{per+07} focused on the relaxation of light objects, stars, by heavy objects, massive perturbers (e.g. giant molecular clouds or clusters, with $M_{H}\\T{\\gg}M_{L}$), and expressed the efficiency of massive perturber-induced relaxation relative to star-star relaxation by the parameter $\\mu_{2}\\T{=}N_{H}M_{H}^{2}/N_{L}M_{L}^{2}$ ($\\sim\\!(D_{HLs}+D_{HLf})/(D_{LLs}+D_{LLf})\\T{\\simeq}3N_{H}M_{H}^{2}/4N_{L}M_{L}^{2}$). When $\\mu_{2}\\T{\\gg}1$, massive perturbers dominate stellar relaxation. Here we focus on the dynamics of the heavy objects, the SBHs, and so the relaxational coupling parameter $\\Delta$ is defined to also take into account the interactions between the heavy masses. The parameter $\\mu_{2}$ addresses the question ``Which mass component dominates the relaxation of the light stars?'', while $\\Delta$ addresses the question ``Which mass component dominates the dynamics of the heavy stars?''. The two are related by $\\Delta\\T{=}\\mu_{2}4/(3+M_{H}/M_{L})$. Our approximate treatment of the mass segregation process neglects the relaxation of angular momentum to near radial ({}``loss-cone'') orbits. This, rather than diffusion in energy, is the primary channel for stellar destruction by the MBH. A full treatment of the mass segregation problem in ($E,J$) phase space \\citep[e.g.][]{coh+78} is beyond the scope of this work. Here we follow BW77, who treated the problem approximately in $E$ only, and who further showed that the neglect of an effective loss-cone term (Eq. \\ref{e:FPeq}) did not much change the shape of the DFs. We confirmed that this conclusion also holds for our mass segregation models (Fig. \\ref{f:Delta_flow}). \\subsection{Possible implications of strong mass segregation} Strong mass segregation occurs in stellar systems with a relatively\\emph{ lower} fraction of SBHs, that reach a \\emph{higher} central concentration of SBHs very close to the MBH, compared to systems with a higher fraction of SBHs that undergo weak segregation. To compare two such systems, and to determine which has more SBHs enclosed inside some given volume around the MBH, it is necessary to specify the comparison procedure (e.g. assuming an equal total stellar mass; or the same MBH mass and the $\\Mbh/\\sigma$ relation; or an equal number of SBHs within $r_{h}$). The choice depends on the question of interest. Here we do not address such quantitative issues, but limit ourselves to briefly listing some processes that are expected to be affected by the degree of segregation. \\emph{Accelerated relaxation. }The degree of mass segregation affects both the non-coherent 2-body relaxation timescale, $ $which scales as $1\\left/\\int M^{2}(\\mathrm{d}N/\\mathrm{d}M)\\mathrm{d}M\\right.$ (see \\S \\ref{s:intro}), and the resonant relaxation timescale, which scales as as $\\int M(\\mathrm{d}N/\\mathrm{d}M)\\mathrm{d}M\\left/\\int M^{2}(\\mathrm{d}N/\\mathrm{d}M)\\mathrm{d}M\\right.$ \\citep{rau+96}. In particular, the stronger the mass segregation, the shorter is the resonant relaxation timescale, which does not depend on the number of stars, but only on their typical mass. Efficient resonant relaxation near the MBH may affect stellar orbits and accretion disk dynamics there (see below). \\emph{GW event rates.} The GW EMRI rate is determined by the number of potential GW sources within the critical radius $r_{\\mathrm{crit}}$, which demarcates the boundary between compact object that inspiral into the MBH those that plunge (infall) into it. The critical radius is a function of the relaxation time, and to good approximation the EMRI rate is $\\Gamma\\T{\\sim}N[r_{crit}(T_{R})]/T_{R}$ \\citep{hop+05}. Strong segregation will affect the EMRI rate both by modifying the 2-body relaxation time and by affecting the number of stars enclosed inside $r_{\\mathrm{crit}}$, as well as by decreasing the resonant relaxation timescale. Similarly, the rates of detectable GW bursts from fly-bys near the Galactic MBH strongly depend on the number of SBHs near it \\citep{hop+07}. Strong segregation may also affect GW emission from close SBH--SBH interactions in a very dense cusp \\citep{ole+08b}. \\emph{SBH--star interactions.} A higher central concentration of SBHs affects the probability of SBH--star interactions, which can lead to the randomization of stellar orbits, the heating of a stellar disk \\citep{per+08}, the 3-body exchange capture of massive young stars near the MBH \\citep{ale+04}, or the ejection of hyper-velocity stars \\citep{ole+08}. \\emph{SBH--accretion disk interactions.} A higher central concentration of SBHs within $10^{1-3}$ gravitational radii of the MBH could exert coherent torques on the accretion disk, warp it and possibly affect its hydrodynamics (Bregman \\& Alexander, 2008, in prep.). The SBHs may shock the disk as they cross it, experience drag by it, and be carried by it to the MBH \\citep[e.g.][]{art+93,nay+04,sub+05}. \\emph{Enhanced gravitational lensing. }SBHs projected near the Einstein angle of the MBH can strongly modify the gravitational lensing properties of the MBH,\\emph{ }in a way similar to the effect of a planet orbiting a Galactic star that is lensing a background source \\citep{ale+01c,cha+01a}. \\subsection{Summary} We show that the steady state solution of a relaxed multi-mass stellar system around a MBH has two branches: the known weak (Bahcall-Wolf) mass segregation solution, where the difference in the degree of central concentration of the light and heavy stars is relatively small, and a newly discovered strong segregation solution, where the difference is much larger. The nature of the solution is determined by the global properties of the system (the mass ratio between the heavy and light stars, $M_{H}/M_{L}$, and their number ratio far from the MBH, $N_{H}/N_{L}$) through the relaxational coupling parameter, $\\Delta\\!=\\!4N_{H}M_{H}^{2}\\left/\\left[N_{L}M_{L}^{2}(3\\!+\\! M_{H}/M_{L})\\right]\\right.$. Strong mass segregation occurs when the heavy stars are relatively rare in the population $(\\Delta\\T{\\ll}1)$, and sink to the center by dynamical friction. Weak mass segregation occurs when the heavy stars are common in the population ($\\Delta\\T{\\gg}1$) and settle to the single mass stellar cusp solution. We show that relaxed old coeval or continuously star-forming populations with a universal IMF typically have $\\Delta\\T{<}0.1$, and thus settle to the strong mass segregation solution around a MBH." }, "0808/0808.1539_arXiv.txt": { "abstract": "{Little is known about the structure of the interstellar medium and the nature of individual clouds in galaxies at intermediate redshifts. The gravitational lens toward PKS\\,1830--211 offers the unique possibility to study this interstellar gas with high sensitivity and angular resolution in a molecular cloud that existed half a Hubble time ago.} {This multi-line study aims at a better definition of the physical properties of a significantly redshifted cloud.} {Using the Green Bank Telescope (GBT), we searched for ammonia (NH$_3$) toward PKS\\,1830--211.} {We have detected all ten observed metastable $\\lambda$$\\sim$2\\,cm ammonia lines. The ($J$,$K$) = (1,1) to (10,10) transitions, up to $\\sim$1030\\,K above the ground state, were measured in absorption against the radio continuum of the lensed background source. The ammonia absorption appears to be optically thin, with absolute peak flux densities up to 2.5\\% of the total continuum flux density. Measured intensities are consistent with a kinetic temperature of $\\sim$80\\,K for 80--90\\% of the ammonia column. The remaining 10--20\\% are warmer, with at least some of this gas reaching kinetic temperatures of $\\ga$600\\,K. Toward the south-western continuum source, the column density is $N$(NH$_3$) $\\sim$ (5--10)$\\times$10$^{14}$\\,cm$^{-2}$, which implies a fractional abundance of $X$(NH$_3$) $\\sim$ (1.5--3.0)$\\times$10$^{-8}$. Similarities with a hot NH$_3$ absorption component toward the Sgr~B2 region close to our Galactic center, observed up to the (18,18) line, suggest that the Sgr~B2 component also consists of warm diffuse low-density gas. The warm absorption features from PKS\\,1830--211 are unique in the sense that they originate in a spiral arm.} {} ", "introduction": "Studies of interstellar emission and absorption lines are complementary. While emission commonly traces an extended gas reservoir, which provides information on global properties such as mass and kinematics, absorption tends to trace a spatially far smaller region confined by the angular extent of a continuum background source. Absorption lines allow us to derive optical depths directly from line to continuum flux density ratios; the strength of the absorption depends on the flux of the background source and is decoupled from the distance to the absorber. Thus, for objects at significant redshifts, studies of interstellar absorption lines have the potential to provide unique information on otherwise inaccessible regions, combining extremely high sensitivity with outstanding spatial resolution (e.g., Chengalur et al. \\cite{chengalur99}; Wiklind \\& Combes \\cite{wc99}). In spite of extended surveys, multi-line molecular absorption systems have so far only been discovered toward five distant targets (Wiklind \\& Combes \\cite{wc94}, \\cite{wc95}, \\cite{wc96a}; \\cite{wc96b}; Kanekar et al. \\cite{kanekar05}), all of which are located at intermediate redshifts (0.25$\\leq$$z$$\\leq$0.89). The two most notable absorbers, those toward B0218+357 and PKS\\,1830--211, share a number of common properties. They have similar redshifts ($z$=0.68 and 0.89, respectively), show the highest line-of-sight column densities, and originate in gravitational lenses. Toward both systems the respective background sources, probably a BL Lac object and a blazar, respectively (e.g. Kemball et al. \\cite{kem01}; De Rosa et al. \\cite{ros05}), are lensed into three dominant features, a north-eastern and a south-western hotspot and an Einstein ring. Optical or near infrared and radio separations of the hotspots differ due to optical obscuration (Courbin et al. \\cite{courbin98}; York et al. \\cite{york05}). The lensing galaxies themselves are spirals viewed almost face-on (Courbin et al. \\cite{courbin02}; Winn et al. \\cite{winn02}; York et al. \\cite{york05}). The bulk of the absorption arises in both cases from in front of the south-western of the two main radio continuum images, at radial distances of $\\sim$2 and $\\sim$4\\,kpc from the center of the respective lens (e.g., Frye et al. \\cite{frye97}; Swift et al. \\cite{swift01}; Meylan et al. \\cite{mey05}; York et al. \\cite{york05}). In both cases, this absorption originates in the south-western spiral arm of the parent galaxy. There are also significant differences between B0218+357 and PKS\\,1830--211: one is image separation (334 versus 970\\,mas; e.g., Jin et al. \\cite{jin03}; Wucknitz et al. \\cite{wuck04}); another is time delay (10 versus 25 days; e.g., Lovell et al. \\cite{lov98}; Biggs et al. \\cite{biggs01}). With respect to the detection of molecular absorption lines, the most striking differences are the continuum levels and the complexity of the absorption systems. B0218+357 hosts two molecular components separated by $\\sim$13\\,km\\,s$^{-1}$ observed toward the south-western continuum image (Menten \\& Reid \\cite{menten96}; Jethava et al. \\cite{jethava07}; Muller et al. \\cite{muller07}). Toward PKS\\,1830--211, weak molecular absorption is also measured toward the north-eastern radio continuum hotspot, displaced by \\hbox{--147}\\,km\\,s$^{-1}$ from the dominant south-western absorption component (Wiklind \\& Combes \\cite{wc98}). Furthermore, there is H{\\sc i} absorption at $z$=0.19 (Lovell et al. \\cite{lovell96}). PKS\\,1830--211 is one of the most prominent compact radio sources in the sky and therefore an ideal target for absorption line studies at radio wavelengths. Molecular species that have been detected include CS, HCN, HCO$^+$, HNC and N$_2$H$^+$ (Wiklind \\& Combes \\cite{wc96b}), CO (Gerin et al. \\cite{gerin97}), OH (Chengalur et al. \\cite{chengalur99}) and C$_2$H, H$_2$CO, C$_3$H$_2$, and HC$_3$N (Menten et al. \\cite{menten99}). The $\\lambda$21\\,cm line of H{\\sc i} was also observed at $z$=0.89 (Chengalur et al. \\cite{chengalur99}). Muller et al. (\\cite{muller06}) determined CNO and sulfur isotope ratios. PKS\\,1830--211 provides a unique view of a molecular cloud at more than half a Hubble time in the past ($\\Lambda$-cosmology with $H_{0}$=73\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_{\\rm m}$ = 0.28 and $\\Omega_{\\Lambda}$ = 0.72; Spergel et al. \\cite{sper07}; 1\\,mas corresponds to $\\sim$7.5\\,pc). With a luminosity distance of $\\sim$5.5\\,Gpc it is the farthest molecular cloud known to date that can be studied in detail. The nature of this object is, however, poorly understood. In an attempt to constrain the physical properties of the target and motivated by the recent detection of NH$_3$ toward B0218+357 (Henkel et al. \\cite{henkel05}), we searched for NH$_3$ absorption to evaluate the still poorly constrained physical parameters of the absorbing molecular gas. ", "conclusions": "Our ammonia observations of the gravitational lens system PKS\\,1830--211 at a redshift of $z$$\\sim$0.9 reveal the following main results: \\begin{itemize} \\item Ammonia (NH$_3$) is detected in absorption in its ten lowest metatstable inversion lines. The ($J$,$K$) = (10,10) transition lies $\\sim$1030\\,K above the ground state. \\item The NH$_3$ absorption lines are optically thin and reach, in the strongest line, only 2.5\\% of the continuum level. No ammonia absorption is detected toward the system displaced by \\hbox{--147}\\,km\\,s$^{-1}$ relative to the main component. \\item The absorbing gas is warm: 80-90\\% of the absorption arises from an environment with $T_{\\rm kin}$ $\\sim$ 80\\,K. The remaining gas is with $>$100\\,K considerably warmer; for some gas, $T_{\\rm kin}$$\\ga$600\\,K. The total column density is approximately $N$(NH$_3$) = (5--10)$\\times$10$^{14}$\\,cm$^{-2}$. \\item Compared with H$_2$ column densities derived from CO at mm-wavelengths, the fractional abundance is $X$(NH$_3$) $\\sim$ 1.5--3.0$\\times$10$^{-8}$. This is a low value when compared with abundances in Galactic cloud cores. The low value may be caused by a low metallicity implying small abundances of nitrogen and dust (NH$_3$ is easily destroyed by UV-photons), by diffuse gas that also shows low NH$_3$ abundances along Galactic lines-of-sight or it may reflect differences between the continuum background morphology at cm- and mm-wavelengths. \\item If the H{\\sc i} spin temperature is equal to the kinetic temperature of the bulk of the NH$_3$ absorbing gas (80\\,K), the total column must be mostly molecular. If, however, most of the H{\\sc i} is associated with the hot ammonia component, the H{\\sc i} and molecular column densities should be similar. \\item In the nearby Universe, only the envelope of Sgr~B2 has a cloud component with ammonia properties similar to those seen toward PKS\\,1830--211. This cloud contains gas from either hot cores or from low-density gas forming an envelope; the analogy with PKS1830--211 suggests the latter. The column toward PKS\\,1830--211 is unique in the sense that it arises from a spiral arm and not from the central region of a gas-rich galaxy. \\end{itemize} Following Flambaum \\& Kozlov (\\cite{flambaum07}), a detailed comparison of NH$_3$ inversion lines with rotational spectra from linear molecules can provide limits on the space-time variation of the ratio of the proton to electron mass. Making use of the ammonia profiles presented here, such a comparison will be carried out in a forthcoming paper." }, "0808/0808.0223_arXiv.txt": { "abstract": "s{Magnetowave induced plasma wakefield acceleration (MPWA) in a relativistic astrophysical outflow has been proposed as a viable mechanism for the acceleration of cosmic particles to ultra high energies. Here we present simulation results that demonstrate the viability of this mechanism. We invoke the high frequency and high speed whistler mode for the driving pulse. The plasma wakefield so induced validates precisely the theoretical prediction. This mechanism is shown capable of accelerating charged particles to ZeV energies in Active Galactic Nuclei (AGN). } ", "introduction": "The origin of ultra high energy cosmic rays (UHECR) is an intriguing question in astrophysics. Theories categorized as either ``top-down\" or ``bottom-up\" scenarios are proposed to answer this question. Each scenario faces its own theoretical and observational challenges \\cite{Olinto}. Since the observations from HiRes \\cite{cosmic:HiRes} and Auger \\cite{cosmic:Auger} confirm the Greisen-Zatepin-Kuz'min (GZK) suppression of the cosmic ray flux \\cite{gzk}, the need for top-down exotic models is reduced. Hence the challenge to find a viable ``bottom up\" mechanism for accelerating UHECR becomes more acute. Shocks, unipolar inductors and magnetic flares are the three most potent, observed, ``conventional\" accelerators that can be extended to account for $\\sim{\\rm ZeV} (=10^{21}{\\rm eV})$ energy cosmic rays \\cite{Blandford:1999}. Radio jet termination shocks and gamma ray bursts (GRB) have been invoked as sites for the shock acceleration, while dormant galactic center black holes and magnetars have been proposed as sites for the unipolar inductor acceleration and the flare acceleration, respectively. Each of these models, however, presents problems \\cite{Blandford:1999}. Evidently, novel acceleration mechanisms that can avoid the difficulties faced by these conventional models should not be overlooked. Plasma wakefield accelerators \\cite{LWFA:Tajima,PWFA:Chen85} are known to possess two salient features: (1) The energy gain per unit distance does not depend (inversely) on the particle's instantaneous energy; (2) The acceleration is linear. These features are essential for a good acceleration efficiency. Although high-intensity, ultra-short photon or particle beam pulses that excite the laboratory plasma wakefields are not available in the astrophysical setting, large amplitude plasma wakefields can instead be excited by the astrophysically abundant plasma ``magnetowaves\" \\cite{plasma:Chen02}. Protons can be accelerated beyond ZeV energy by riding on such wakefields. This attractive concept has never been validated through self-consistent computer simulations. In this presentation, we report our simulation that confirm this concept \\cite{Chang:2007um}. We also discuss this acceleration mechanism in AGN. ", "conclusions": "Through PIC simulations, we have confirmed the concept of plasma wakefield excited by a magnetowave in the magnetized plasma. We have demonstrated how such a wakefield may accelerate particles to ZeV energies in AGN. As a first step, we investigated MPWA in the parallel-field configuration. Since both poloidal and toroidal field components are inevitable in astro-jets, we will further investigate plasma wakefield excitation and acceleration under the cross-field configuration. Besides, we have limited our discussions in the linear regime $a_0\\ll 1$, which is applicable to AGN. However, in other astrophysical settings, such as GRB, the magnetowave could be much stronger such that $a_0\\gg 1$. We will investigate plasma wakefield generations in this regime and therefore explore the MWPA production of UHECR in other powerful astrophysical sites." }, "0808/0808.0153_arXiv.txt": { "abstract": "We present deep HST ACS/WFC photometry of the dwarf irregular galaxy NGC 1569, one of the closest and strongest nearby starburst galaxies. These data allow us, for the first time, to unequivocally detect the tip of the red giant branch and thereby determine the distance to NGC 1569. We find that this galaxy is 3.36$\\pm$0.20 Mpc away, considerably farther away than the typically assumed distance of 2.2$\\pm$0.6 Mpc. Previously thought to be an isolated galaxy due to its shorter distance, our new distance firmly establishes NGC 1569 as a member of the IC 342 group of galaxies. The higher density environment may help explain the starburst nature of NGC 1569, since starbursts are often triggered by galaxy interactions. On the other hand, the longer distance implies that NGC 1569 is an even more extreme starburst galaxy than previously believed. Previous estimates of the rate of star formation for stars younger than $\\la$ 1 Gyr become stronger by more than a factor of 2. Stars older than this were not constrained by previous studies. The dynamical masses of NGC 1569's three super star clusters, which are already known as some of the most massive ever discovered, increase by $\\sim$53\\% to 6-7$\\times10^5$\\msun. ", "introduction": "Massive starbursts drive the evolution of galaxies at high redshift; they provide chemical enrichment and thermal and mechanical heating of both the interstellar and intergalactic medium. In the past decade, a large population of star-forming galaxies has been discovered at high redshift, highlighting the importance of these galaxies on a cosmological scale. However, starburst galaxies can only be studied in detail in the nearby Universe where they are much rarer. The dwarf irregular galaxy NGC 1569 is one of the closest examples of a true starburst. Star formation in NGC 1569 has been studied extensively with the Hubble Space Telescope (HST; e.g., Greggio et al.~1998, hereafter G98, Aloisi et al.~2001, Angeretti et al.~2005, hereafter A05), with results showing that its star formation rates (SFRs) are 2-3 times higher than in other strong starbursts (e.g.~NGC 1705) and 2-3 orders of magnitude higher than in Local Group irregulars and the solar neighborhood, if the SFR per unit area is considered. NGC 1569 is also home to three of the most massive super star clusters (SSCs) ever discovered.\\looseness=-2 Although it lies on the sky in the same direction as the IC 342 group of galaxies (see Fig.~2 in Karachentsev et al.~2003), the typically assumed distance of 2.2$\\pm$0.6 Mpc (Israel 1988) is based on the luminosity of the brightest resolved stars (Ables 1971) and places NGC 1569 well in front of IC 342 ($D = 3.28 \\pm 0.27$ Mpc; Saha, Claver, \\& Hoessel 2002). NGC 1569 has therefore generally been viewed as an isolated starburst galaxy just beyond the outskirts of the Local Group. The isolated environment makes it more complicated to understand the trigger of the starburst, since starbursts are often associated with galaxy interactions and are not generally believed to be internally driven. However, NGC 1569's distance is relatively uncertain, and a large range of possible distances exists in the literature. For example, using HST/WFPC2 photometry, Makarova \\& Karachentsev (2003) attempted to measure the brightness of stars at the tip of the red giant branch (TRGB) in NGC 1569 and found, due to the limited number of stars available at their detection limit, two possible solutions; a short distance of $1.95\\pm0.2$ Mpc or a long distance of $2.8\\pm0.2$ Mpc, with no preference for one over the other. We note also that distance determinations based on the brightest resolved stars in dwarf galaxies are prone to considerable uncertainty due to stochastic effects (e.g.~Greggio 1986), and that, at the distance of NGC 1569, ground-based observations suffer from the possibility of misinterpreting star clusters or pairs of stars as individual stars. In an effort to better determine its star formation history (SFH), we have used the HST ACS/WFC to obtain deep $V$- and $I$-band photometry of NGC 1569. These data also provide for the first time an accurate distance based on unequivocal identification of the TRGB, which we present and discuss in this Letter. ", "conclusions": "" }, "0808/0808.2156_arXiv.txt": { "abstract": "A new method for determining the stellar rotation period is proposed here, based on the detection of starspots during transits of an extra-solar planet orbiting its host star. As the planet eclipses the star, it may pass in front of a starspot which will then make itself known through small flux variations in the transit light curve. If we are lucky enough to catch the same spot on two consecutive transits, it is possible to estimate the stellar rotational period. This method is successfully tested on transit simulations on the Sun yielding the correct value for the solar period. By detecting two starspots on more than one transit of HD 209458 observed by the Hubble Space Telescope, it was possible to estimate a period of either 9.9 or 11.4 days for the star, depending on which spot is responsible for the signature in the light curve a few transits later. Comparison with period estimates of HD209458 reported in the literature indicates that 11.4 days is the most likely stellar rotation period. ", "introduction": "Four centuries have elapsed since the first determination of the rotational period of a star, namely the Sun, by observation of the apparent movement of starspots on its surface. For star other than the Sun, the determination of the rotational period is made basically by two methods: either from the rotational broadening of spectral lines or by the periodic modulation of the stellar flux due to the dark and bright features on the stellar surface which rotate with it. The latter method has the advantage of determining directly the period of a star, without the sin$i$ spectroscopic uncertainty. Moreover, it can also be applied to stars with long rotational period, which cannot be determined from Doppler broadening of spectral lines. Here I would like to propose a new way of estimating the stellar rotation by using planetary transits. If by any chance, during a transit, the planet passes in front of a starspot, as in Silva (2003), and in a consecutive transit the configuration is such that this same spot is again occulted, then it is possible to estimate the stellar rotation as it was done for the Sun four centuries ago. Previous authors have used a similar method to estimate, among other things, the rotational period of the components of eclipsing binaries. Eaton and Hall (1979) explained successfully the light curve variations of RS CVn type stars as due to starspots. Besides the rotation period and its variation, durations of activity cycles and surface area coverage of spots have also been obtained for the prototypes of active binaries RS CVn (Rodono et al., 1995) and RT Lac (Lanza et al. 2002). These are very active stars, nevertheless it will be possible to study moderately active stars such as our Sun with space missions such as CoRoT, MOST, and Kepler. In anticipation to these satellites, the rotational modulation of the Sun as a star has been modeled using the VIRGO/SoHO data (Lanza et al. 2003, 2004). The method described here, however, has a much better spatial resolution than that of eclipsing binaries, because of the smaller planet diameter in comparison to that of the stars. Moreover, planetary transits such as the ones used here will be quite commonly detected by the CoRoT mission already in operation. Section \\ref{transit} presents the method, which is tested on the Sun with a fiducially transiting planet (Section \\ref{solar}) and then applied to HD209458 (Section \\ref{hd}). The last section discusses the results of this method and presents the conclusions. ", "conclusions": "This work proposes to estimate the rotational period of a star by following the apparent shift in longitude position of its surface spots, similar to what was done by Galileo and his contemporaries four centuries ago for the Sun. Tracking of the spots is done by identifying ``bumps\" in the light curves of successive planetary transits. This method was successfully tested for the Sun yielding the correct value for the solar period from simulated transits taken only three days apart even though the solar period is about 27 days. Moreover, by modeling of different sunspot groups it was possible to verify the solar differential rotation. Supposing that the small variations detected in the light curve during planetary transits is due to occultation of starspots, the model was also applied to HD 209458 using HST observations obtained by Brown et al. (2001) in April and May of 2000. The star was modeled as having two spots during the four transits. The ``bumps\" detected on two transits separated by three orbital periods yield a stellar rotation period of 9.9 or 11.4 days depending on which of the spots detected on April 25th is considered to cause the intensity variation on the May 5th transit light curve. Several observations of $v \\sin i$ of HD 209458 are found in the literature. Assuming that the inclination angle is 86.68$^\\circ$ and the stellar radius 1.148 R$_\\odot$, these observations yield stellar periods of 14.4 $\\pm$ 2.1 days (Mazeh et al. 2000) and 15 $\\pm$ 6 days (Queloz et al. 2000). More recently, shorter periods have been found, 12.3 $\\pm$ 0.5 days (Winn et al. 2005) and 12 days (Fisher \\& Valenti 2005). The periods obtained here, 9.9 and 11.4 days, are a little shorter than those found in the literature. It seems that the 11.4 days period is the true one for HD 209458, even though it is still a little shorter than the periods listed in the literature. However, HD 209458 probably presents differential rotation, similar to the Sun. In this case, the period obtained by other authors, which was based on line broadening observations, is actually an average of the periods of the whole stellar disk, whereas the period determined here represents the rotational velocity at that specific latitude which is close to the equator. As the planet size spans about 20$^o$ in stellar latitude, we are probing latitudes from -22$^o$ to -38$^o$. Since the work of Winn et al. (2005) was based on the observation of the Rossiter-McLaughlin effect during transits, therefore obtained at the same latitudes we sample, this is the result which should better agree with the one presented here. Thus far, over 50 transiting planets have been detected (The Extrasolar Planets Encyclopaedia - http://exoplanet.eu) and many more are expected especially in the next months thanks to observations by the CoRoT satellite. The method proposed here can easily be applied to the planetary transits of newly discovered planets and checked against periodic modulation of the stellar flux outside transits themselves." }, "0808/0808.3293_arXiv.txt": { "abstract": "Tidal tails of star clusters are not homogeneous but show well defined clumps in observations as well as in numerical simulations. Recently an epicyclic theory for the formation of these clumps was presented. A quantitative analysis was still missing. We present a quantitative derivation of the angular momentum and energy distribution of escaping stars from a star cluster in the tidal field of the Milky Way and derive the connection to the position and width of the clumps. For the numerical realization we use star-by-star $N$-body simulations. We find a very good agreement of theory and models. We show that the radial offset of the tidal arms scales with the tidal radius, which is a function of cluster mass and the rotation curve at the cluster orbit. The mean radial offset is 2.77 times the tidal radius in the outer disc. Near the Galactic centre the circumstances are more complicated, but to lowest order the theory still applies. We have also measured the Jacobi energy distribution of bound stars and showed that there is a large fraction of stars (about 35\\%) above the critical Jacobi energy at all times, which can potentially leave the cluster. This is a hint that the mass loss is dominated by a self-regulating process of increasing Jacobi energy due to the weakening of the potential well of the star cluster, which is induced by the mass loss itself. ", "introduction": "\\label{sec-intro} Recently well-defined clumps were observed in the tidal tails of globular clusters by \\citet{Le00} for NGC 6254 and Pal 12 and by \\citet{Od01,Od03} for Pal 5. External perturbations like crossing of the galactic disc, peri-centre passage or near-encounters with other globular clusters were discussed as the source of the clumps \\citep{Ca05}. In \\citet{Ca05} the formation of these clumps in tidal tails of star clusters on eccentric orbits were confirmed by numerical studies. But the clumps occur also in the tidal tails of star clusters on circular orbits with no external push. Recently, \\citet{Ku08} presented a theoretical explanation for the clump formation in a constant tidal field. It is essentially due to the epicyclic motion of the stars lost by the star cluster. We present a quantitative analysis of the tidal tail structure for star clusters moving on a circular orbit in the galactic disc. The analysis is based on numerical simulations with realistic particle numbers including an initial mass function (IMF) and stellar evolution. We compare the results for star clusters at the solar circle and near the Galactic centre. Since the epicycle theory is a perturbation theory with respect to a circular orbit with constant tidal field, we cannot apply it for predictions of clump distances to the eccentric orbits of the observed globular clusters. On the other hand the observation of tidal tail clumps of open clusters in the galactic disc are hampered by the overwhelming number of field stars with similar properties. For an identification the contrast in density and velocity with respect to the field stars may be too small. Additionally the tidal tails may be destroyed quickly by the same gravitational scattering process, which is also responsible for the dynamical heating of the stellar disc. We discuss the observability further in Sect.~\\ref{sec-sum}. In Section \\ref{sec-dyn} we present the epicyclic theory for the stars in the tidal tails and the connection to the mass loss and the orbit of the star cluster. In Section \\ref{sec-num} we present the numerical codes used and the properties of the star clusters. Section \\ref{sec-res} contains the quantitative comparison of the numerical results with the theoretical predictions. In Section \\ref{sec-sum} we summarize our results. ", "conclusions": "\\label{sec-sum} We presented a quantitative derivation of the angular momentum and energy distribution of escaping stars from a star cluster in the tidal field of the Milky Way. Despite the motion on a circular orbit, the tidal tails are clumpy due to the epicyclic motion of the stars. We compared the derived distances and widths of the clumps with numerical simulations using star-by-star simulations. For star clusters at the solar circle we included an IMF and mass loss due to stellar evolution in the calculations. The same equations were applied to a star cluster very close to the Galactic centre, where the tidal forces are very strong. We find a very good agreement of theory and models concerning the tidal tail structure. The positions of the clumps are determined by the angular momentum offset of the stars, which lead to a radial offset of the epicenters with respect to the cluster orbit. The investigation of \\citet{Ku08} is a special case of our investigations but for a Kepler potential. We find that the radial offset of the tidal arms is proportional to the tidal radius. However near the Galactic centre the factor of proportionality is considerably smaller. The tidal arm structure at large galactocentric radii is symmetric, whereas the asymmetry near the galactic centre is considerable. This can be reproduced only partly by taking into account the correction of the epicyclic frequency at the epicentre radii. We have also measured the Jacobi energy distribution of bound stars and showed that there are 35\\% of stars above the critical Jacobi energy independent of the evolutionary state of the cluster. These stars can potentially leave the cluster. This is a hint, that mass loss is dominated by a self-regulating process of increasing Jacobi energy due to the diminishing gravitational potential of the star cluster induced by the mass loss itself. Finally we consider the observability of the predicted clump properties in the tidal tails of star clusters. The identification of tidal tail stars of open clusters on a circular orbit is strongly hampered by the large number of nearby field stars with similar properties. But with differential methods using high quality data for distances and velocities it may be possible to identify the most prominent first clumps. We have shown that the maximum density in the first clump does not decrease with time until dissolution of the cluster. The first clumps are formed by escaping stars with a time delay determined by the epicyclic period $T_\\kappa\\approx 150$\\,Myr. The tidal tail structure will probably survive the gravitational scattering process, which is also responsible for the galactic disc heating \\citep{Wie77}. On a timescale of one epicyclic period, we expect only small perturbations of the tidal clump position and velocity but no destruction. The density maximum of the first clump is of the order of a few percent of the field density of the galactic disc. The velocity imprint by the epicyclic motion is of the order of 2\\,km/s. An overdensity with these properties may be difficult to observe for clusters on exact circular orbits, if there is no additional separating property. For young star clusters the high fraction of early type stars can serve for such a discrimination. On the other hand most star clusters are identified by the systematic peculiar motion with respect to the field stars. If the positions and velocities of the tidal clump stars are properly predicted, they may be observable as moving groups. For an application of the tidal clump theory to the eccentric orbits of globular clusters a perturbation theory with respect to the 'free falling' comoving coordinate system would be necessary. In this case the Jacobi energy is no longer a constant of motion. This will be a matter of future investigations. Some numerical test runs have shown that tidal tail clumps are formed also on highly eccentric orbits without an additional external perturbation, but the geometry and density is modulated along the orbital position." }, "0808/0808.1069_arXiv.txt": { "abstract": "{The remarkable pre-main-sequence object \\mbox{V718 Per} (HMW 15, H187) in the young cluster IC 348 periodically undergoes long-lasting eclipses caused by variable amounts of circumstellar dust in the line-of-sight to the star. It has been speculated that the star is a close binary and similar to another unusual eclipsing object, KH 15D.} {We have submitted \\mbox{V718 Per} to a detailed photometric and spectroscopic study to investigate how regular the recurrent eclipses are, to find out more about the properties of the stellar object and the occulting circumstellar material, and to look for signatures of a possible binary component.} {\\mbox{V718 Per} was monitored photometrically from the optical to the near-infrared (NIR). We also obtained high-resolution optical spectra with the Keck telescope at minimum as well as at maximum brightness. We derived the fundamental photospheric parameters of this star by comparing with synthetic spectra.} {Our photometric data show that the eclipses are very symmetric and persistent, and that the extinction law of the foreground occulting dust deviates only little from what is expected for ``normal'' interstellar material. The stellar parameters of \\mbox{V718 Per} indicate a primordial abundance of Li and a surface temperature of $T_{\\rm eff} \\approx 5200$ K. Remarkably, the in-eclipse spectrum shows a significant broadening of the photospheric absorption lines, as well as a slightly lower stellar surface temperature. In addition, weak emission components appear in the absorption lines of H$\\alpha$ and the Ca II IR triplet lines. We did not detect any signs of atomic or molecular features related to the occulting body in the in-eclipse spectrum. We also found no evidence of any radial velocity changes in \\mbox{V718 Per} to within about $\\pm 80$ m s$^{-1}$, which for an edge-on system corresponds to a maximum companion mass of ${\\sim}6~M_{\\rm Jup}$.} {Our observations suggest that \\mbox{V718 Per} is a single star, and thus very different from the related binary system KH 15D. We conclude that \\mbox{V718 Per} is surrounded by an edge-on circumstellar disk with an irregular mass distribution orbiting at a distance of 3.3 AU from the star, presumably at the inner disk edge. To produce the prolonged eclipses, the occulting feature must extend along more than half of the inner disk edge. The change in stellar surface temperature and the emission line activity observed could be related to spot activity. We ascribe the broadening of photospheric absorption lines during the eclipse to forward scattering of stellar light in the circumstellar dust feature.} ", "introduction": " ", "conclusions": "Our new photometric data of \\mbox{V718 Per} (H187) extends previous measurements and confirms that this object shows long-lasting eclipses with a period of 4.7 years. The eclipses, which are very symmetric, are caused by occulting dust, and the colour changes suggest an extinction law of the foreground dust that deviates only little from what is expected for ``normal'' interstellar grains. It has been speculated that \\mbox{V718 Per} may be similar to the unusual close binary KH 15D, also showing periodic eclipses, which presumably are caused by different coverage of the orbiting stars by a circumbinary disk. We obtained two high-dispersion spectra of \\mbox{V718 Per}, the first close to the deepest point of the eclipse and the other at a time outside eclipse with a time difference corresponding to roughly a quarter of the eclipse period. From these spectra we found no evidence of any change in the radial velocity of the star to within $\\pm 80$ m s$^{-1}$. Although we cannot be fully certain on the basis of two spectra, the absence of radial velocity variations in a system that experiences eclipses from its circumstellar material makes it very unlikely that \\mbox{V718 Per} is a close binary. \\mbox{V718 Per} seems therefore very different from KH 15D. We confirm that the eclipses are caused by a stable extended, dusty structure orbiting at ${\\sim} 3.3$ AU from the star in an edge-on circumstellar disk. We have derived the SED of \\mbox{V718 Per} outside eclipse, including IR fluxes obtained with {\\it Spitzer}. This indicates that the star could be surrounded by a thin, low-mass disk. Because of the extended eclipse duration, the structure that causes the eclipses must extend over half a circle along the disk edge. A low-mass companion can in principle induce this structure, but the mass of any planet cannot exceed $6~M_{\\rm Jup}$ (assuming an edge-on system). The full amplitude of an eclipse amounts to 1.1 magnitudes in the V-band, but there are no signs of any enhanced absorption features from circumstellar gas during eclipse. It appears that the occulting structure is rather void of gas. \\mbox{V718 Per} has a typical late-type absorption line spectrum without strong emission lines of e.g H$\\alpha$. We have used theoretical synthetic spectra from model atmosphere and derived fundamental photospheric parameters. Our spectroscopic analysis shows that \\mbox{V718 Per} has a primordial abundance of Li and a surface temperature of $T_{\\rm eff} \\approx 5200$\\,K. From the luminosity derived from its SED, and by comparing with theoretical evolution tracks, we find that \\mbox{V718 Per} is in its post-T Tauri phase of evolution. However, there are remarkable differences between the in-eclipse and out-of-eclipse spectrum. During the eclipse several spectral features show that the surface temperature is slightly lower than in the out-of-eclipse spectrum, corresponding to a change in spectral type from G9 to K0. In addition, narrow emission components appear in the absorption cores of the H$\\alpha$ and the Ca {\\sc ii} IR triplet lines during eclipse, and the photospheric absorption lines become slightly broader. The change in stellar surface temperature and the emission line activity observed is puzzling. {\\it Since \\mbox{V718 Per} shows no short-term (rotational) photometric variability, this cannot be explained as the result of a variable coverage by starspots.} However, the observed spectral changes could be related to long-term changes in the activity near the polar regions. The broadening of the absorption lines we ascribe to forward scattering of stellar light in the circumstellar dust feature when it passes through the line-of-sight." }, "0808/0808.1633_arXiv.txt": { "abstract": "By using images taken with WFCAM on UKIRT and SofI on the NTT and combining them with 2MASS we have measured proper motions for 126 L and T dwarfs in the dwarf archive. Two of these L dwarfs appear to have M dwarf common proper motion companions, and 2 also appear to be high velocity dwarfs, indicating possible membership of the thick disc. We have also compared the motion of these 126 objects to that of numerous moving groups, and have identified new members of the Hyades, Ursa Major and Pleiades moving groups. These new objects, as well as those identified in \\citet{jameson08} have allowed us to refine the L dwarf sequence for Ursa Major that was defined by \\citet{jameson08b}. ", "introduction": "Brown dwarfs may be thought of as failed stars. These low mass ($\\leq$70 M$_{\\rm Jup}$ \\citealt{burrows01}), cool objects are the lowest mass objects that the star formation process can produce. The majority of the brown dwarfs that have been discovered to date are field objects found using surveys such as the Two Micron All Sky Survey (2MASS; \\citealt{skrutskie06}, see \\citealt{leggett02} for examples), the DEep Near-Infrared Sky survey (DENIS; \\citealt{denis05}, see \\citealt{delfosse99} for examples), the Sloan Digital Sky Survey (SDSS;\\citealt{york00} see \\citealt{hawley02} for examples) and the UKIRT Deep Infrared Sky Survey (UKIDSS; \\citealt{lawrence06}, see \\citealt{kendall07, lodieu07b} for examples). However, to study brown dwarfs in depth, a knowledge of their age is essential, which means we must study brown dwarfs in open star clusters or moving groups. Once a brown dwarf has been proved to belong to an open star cluster, or a moving group, then its age is known, allowing meaningful comparisons to evolutionary models to be made. The most recent example of this is the study done by \\citet{bannister07} who used existing proper motions and parallax measurements to show that a selection of field dwarfs in fact belong to the Ursa Major and Hyades moving groups. The importance of this study, is that these are the first brown dwarfs to be associated with an older cluster or group (age $>$200 Myr). Older clusters such as the Hyades are expected to contain very few or no brown dwarfs or low mass members, due to the dynamical evolution of the cluster over time \\citep{adams02}. However, these escaped low mass objects may remain members of the much larger moving group that surrounds the cluster. \\citet{jameson08} followed this work by using the wide field camera (WFCAM, \\citealt{casali07}) on the United Kingdom Infrared Telescope (UKIRT) to image 143 known field L dwarfs. These images provided a second epoch for proper motion measurements, when combined with existing 2MASS images, typically taken 7 years previously. Using the proper motions and a distance calculated using the spectral type of the L dwarf given by \\citet{cruz03}, the moving group method was applied, and all 143 objects were scrutinised to check if their direction and magnitude of motion made them candidates of the many moving groups known. Members of the Hyades, Ursa Major and Pleiades moving groups were found. Radial velocity measurements such as those of \\citet{zapatero07} are required however, before it can be determined if these moving group members are cluster members that have ``escaped'' as the cluster has dynamically evolved \\citep{adams02}. Is should also be noted that galactic resonances can produce effects similar to moving groups, and so all members may not be coeval \\citep{dehnen98}. To continue the study started by \\citet{bannister07} and \\citet{jameson08}, we have measured proper motions for the majority of the remaining known field L dwarfs listed in the online L and T dwarf archive (http://spider.ipac.caltech.edu/staff/davy/ARCHIVE/). This has again been accomplished using the WFCAM on UKIRT and for the more southern objects, Son of ISAAC (SofI) on the 3.58m ESO New Technology Telescope (NTT). Using these wide field images and existing catalogue data, we have measured proper motions for an additional 126 L and T dwarfs listed in the dwarf archive. These proper motion data may be put to a number of uses. Using reduced proper motion diagrams they can be used as an approximate measure of distance. The proper motion measurements can also be used to help identify objects as members of a star cluster or members of a moving group via the moving cluster method. Taken with measured radial velocities and distances, it can yield all three components of velocity (U,V,W). As brown dwarfs tend to be faint, measuring their radial velocities is very difficult. As a result, very few L or T dwarfs have measurements, none of which are in this sample. This means we cannot determine whether these moving group members are escaped cluster members or otherwise. Our proper motion data are discussed and listed in section 2 of this paper. ", "conclusions": "" }, "0808/0808.3400_arXiv.txt": { "abstract": "We present a framework for analyzing weak gravitational lensing survey data, including lensing and source-density observables, plus spectroscopic redshift calibration data. All two-point observables are predicted in terms of parameters of a perturbed Robertson-Walker metric, making the framework independent of the models for gravity, dark energy, or galaxy properties. For Gaussian fluctuations the 2-point model determines the survey likelihood function and allows Fisher-matrix forecasting. The framework includes nuisance terms for the major systematic errors: shear measurement errors, magnification bias and redshift calibration errors, intrinsic galaxy alignments, and inaccurate theoretical predictions. We propose flexible parameterizations of the many nuisance parameters related to galaxy bias and intrinsic alignment. For the first time we can integrate many different observables and systematic errors into a single analysis. As a first application of this framework, we demonstrate that: uncertainties in power-spectrum theory cause very minor degradation to cosmological information content; nearly all useful information (excepting baryon oscillations) is extracted with $\\approx 3$ bins per decade of angular scale; and the rate at which galaxy bias varies with redshift substantially influences the strength of cosmological inference. The framework will permit careful study of the interplay between numerous observables, systematic errors, and spectroscopic calibration data for large weak-lensing surveys. ", "introduction": "Weak gravitational lensing of background sources can produce exceptionally strong constraints on cosmological parameters and tests of General Relativity. Initial analyses considered the 2-point correlation function (or, equivalently, power spectrum) of the shear pattern induced on a single population of background galaxies \\citep{Jordi, Kaiser92, Blandford91}. A wealth of new statistics, however, have been suggested as more powerful means to extract information from weak lensing (WL): cross-power spectra of multiple source populations with distinct redshift distributions (a.k.a. ``tomography'') \\citep{Hu}; the correlation of shear with foreground galaxy clusters \\citep{JainTaylor}, or more generally the cross-correlation of lensing shear with the galaxy distribution \\citep{BJ04, Zhang}; joint analyses of density-density, density-shear, and shear-shear correlations in an imaging survey \\citep{HJ04}; cross-correlation of magnification as well as shear \\citep{Jainmag}; use of the CMB \\citep{HO02,HS03} or recombination-era 21~cm signals \\citep{MW07,Pen04,ZZ06} as source planes; cross-correlation of source density or shear with a distinct spectroscopic galaxy survey population \\citep{Newman, SKZC}; and the use of 3-point statistics \\citep{TJ04} or statistics such as peak counts \\citep{Hennawi,Wang,Laura} to move beyond 2-point information. Each of these potential innovations has been individually analyzed and shown to improve cosmological constraints. The first goal of this paper is to consider the {\\em simultaneous} use of all of these observable statistics: can we forecast the cosmological information that they will yield collectively in future surveys? Can we start to develop a framework in which all these signals could be analyzed simultaneously in a real experiment? In parallel with the increasing variety of proposed WL signals, the community has identified a series of potential astrophysical and instrumental non-idealities in WL data which, if ignored, would lead to substantial systematic errors in the inferred cosmology. These include: finite accuracy in our ability to predict the deflecting mass power spectrum due to nonlinearities \\citep{JS97} and baryonic physics \\citep{Zhan06, Jing}; intrinsic alignments (IA) between galaxy shapes \\citep{CroftMetzler} and between galaxy shapes and the local mass distribution \\citep{Hirata2} that are not induced by lensing; multiplicative ``shear calibration'' errors in the derivation of lensing shear from galaxy images \\citep{Ishak04,HTBJ}; additive ``spurious shear'' due to uncorrected PSF ellipticity or other imaging systematics \\citep{HTBJ,AmaraRefregier}; and errors in the assignment of redshifts to the source populations \\citep{MaHutererHu}. The impact of these systematic-error sources on cosmological inferences have been analyzed by different means, but a second goal of this paper is to produce a comprehensive forecast that considers the presence of them all simultaneously. Previous work has shown that these multiple sources of information and systematic in WL surveys can interact in interesting ways. For example, in the presence of tomographic data, many systematics are readily distinguishable from cosmological signals and can hence be diagnosed and corrected internally to a survey; this approach is called {\\em self-calibration} \\citep{HTBJ}. It has also been shown that combining galaxy density and lensing correlations can lead to self-calibration of shear calibration errors \\citep{BJ04} and the uncertainties in galaxy biasing \\citep{HJ04, Zhan06}. Intrinsic alignments of galaxies can be diagnosed and corrected if tomographic information is available \\citep{KingSchneider}, however this places substantially greater demands on the precision and accuracy of redshift assignment than would otherwise be needed \\citep{BridleKing}. These investigations raise important practical questions: will the self-calibration techniques continue to succeed when we attempt to simultaneously self-calibrate several different systematic errors? Do cross-correlation techniques reduce uncertainties in redshift distributions to negligible levels, or is it necessary to make a complete spectroscopic redshift survey of some size to measure redshift distributions directly \\citep{MaBernstein}? This paper will present a formalism through which all these questions can be answered, but we defer to later papers the application of the framework to these issues. A third goal of this work is to describe the constraints by WL in a language that is not tied to a specific cosmological model. Most forecasts for WL survey constraints are done within the context of a Universe that has homogeneous dark energy with equation of state $w=w_0 + w_a(1-a)$. Projecting the WL experiment onto this model gives concrete predictions, but obscures what the WL is really measuring. So the analysis framework presented here will be {\\em dark-energy agnostic}, meaning that no specific model is assumed. We will be very explicit about the assumptions made in the analysis and try to keep them to a minimum. In fact a great strength of WL experiments are their ability to test General Relativity itself, so we seek an analysis method that is general enough to incorporate such tests. Similar to the approach of \\citet{KST}, our analysis results in constraints on the distance and growth functions $D(z)$ and $g_\\phi(z)$, without reference to the particular dark-energy or gravity modifications that might cause deviations from $\\Lambda$CDM. In the following section we describe a ``kitchen-sink'' formalism for WL survey observables that allows the incorporation of all suggested 2-point statistics and very general treatments of nearly all proposed systematic errors. In \\S\\ref{speclike} we give a likelihood function and Fisher matrix for an unbiased spectroscopic redshift survey of source galaxies. Then we briefly describe a software implementation of the lensing and spectroscopy likelihood calculations. We describe our model for the evolution of the lensing-potential power spectrum in \\S\\ref{powermodel}, and \\S\\ref{nuisance} we describe generic models used for the nuisance functions required in the lensing-survey analysis. In \\S\\ref{tuneup} we use the implementation of these methods to investigate the proper choices for the bin sizes and grid spacings needed to turn the lensing analysis into a tractable finite-dimensional problem. Further application of the framework to survey forecasting will be done in future papers. An earlier version of this WL analysis formalism was used to generate forecasts for the Dark Energy Task Force \\citep{DETF}, and is described in an appendix to that report. ", "conclusions": "The core of this paper are the expressions (\\ref{gg1})--(\\ref{kk1}) for the two-point correlation matrix of the lensing and density observable multipoles produced by a typical lensing survey. This was derived under a very limited set of assumptions: a homogeneous and isotropic 4-dimensional metric Universe with scalar perturbations; plus the weak-lensing limit, the Limber and Born approximations, and an approximation that lensing magnification bias and intrinsic density fluctuations are additive. The last four assumptions could be relaxed at the expense of computational complexity. We thus hope that data analyses based on this framework could be used to constrain a wide variety of potential explanations for the acceleration phenomenon, including gravity modifications as well as new fields in the Universe. In the limit of Gaussian fluctuation fields, the two-point information is a complete description of the likelihood and hence can be used to construct Fisher matrices or analyze data. As currently configured, the analysis yields the survey's ability to constrain the distance function $D(z)$ and linear growth function $g_\\phi(z)$, without reference to particular dark-energy models. It would be straightforward to implement scale-dependent linear-growth functions. This framework subsumes all of the information (up to 2-point level) that is likely to be obtained from lensing observations: density-density, lensing-density, and lensing-lensing correlations, plus redshift distributions from unbiased spectroscopic surveys (\\S\\ref{speclike}). Furthermore it allows for the most important expected forms of systematic error: photo-z calibration errors, shear and magnification-bias calibration errors, intrinsic alignments, and inaccuracies in power-spectrum theory. Systematics that are additive to shear (\\eg uncorrected PSF ellipticity) or to density (\\eg uncorrected foreground extinction) have not been included. We have not done so since the additive errors could, in principle, exhibit almost any arbitrary signature in the covariance matrix of the observables. Hence a completely general model for additive errors would be degenerate with almost all other signals. For the additive systematics, it is better to determine the level at which they would bias the cosmological results than to attempt to fit a model. \\citet{AmaraRefregier} is a good example of this approach. Since the analysis framework is independent of models for dark energy, gravity, power-spectrum evolution, or galaxy bias, we get a stripped-down look at what parameters are truly constrained by the data, and what nuisance functions must be modeled in order to extract the cosmological information. There is a substantial suite of biases and correlation functions involved in understanding the full survey data. In other work these have been ignored, or have been quantified by reference to halo occupation models \\citep{HJ04,CvdBMLMY}. Here we introduce generic functions for bias and calibration nuisance functions that are not based on any particular physical model. We implement one possible model for the evolution of the lensing-potential power spectrum, based on General Relativity but allowing for failure of the growth equation. It is straightforward to implement other potential deviations from General Relativity. In the current implementation, the end result of the Fisher analysis is a forecast of the ability to constrain the functions $D_A(z)$ and $g_\\phi(z)$. Since the analysis must be discretized in redshift and angular scale in order to be feasible, we investigated the bin sizes or bandwidths of nuisance functions that should be chosen. We find that $\\approx 3$ bins per decade of angular scale suffice to extract all information (apart from baryon acoustic oscillations), and that nuisance functions should be specified no finer than this. Nuisance functions for power-spectrum theory errors and for shear and magnification-bias calibration errors can be specified coarsely in redshift space ($\\Delta \\ln a \\approx 1$), but the galaxy biases, correlations, and intrinsic-alignments must be modeled with potentially finer structure in redshift ($\\Delta \\ln a \\approx 0.1$) to immunize against potential astrophysical systematics. In future papers we will use this framework and its implementation to investigate the requirements for spectroscopic calibration of photo-z's in large lensing surveys, and other practical issues. As a simple first application of our framework, we have shown here that power-spectrum theory uncertainty does not significantly degrade the cosmological power of a nominal lensing survey at $10<\\ell<10^4$. Non-Gaussian statistics are a much more important factor to consider. {\\tt C++} Code to implement Fisher forecasting using this framework has been written and runs quickly on desktop computers despite the large number of free parameters in these general models. Interested parties should contact the author for access to the code." }, "0808/0808.0129_arXiv.txt": { "abstract": "Tightest known quadruple systems VW~LMi consists of contact eclipsing binary with $P_{12}$ = 0.477551 days and detached binary with $P_{34}$ = 7.93063 days revolving in rather tight, 355.0-days orbit. This paper presents new photometric and spectroscopic observations yielding 69 times of minima and 36 disentangled radial velocities for the component stars. All available radial velocities and minima times are combined to better characterize the orbits and to derive absolute parameters of components. The total mass of the quadruple system was estimated at 4.56 M$_\\odot$. The detached, non-eclipsing binary with orbital period $P$ = 7.93 days is found to show apsidal motion with $U \\approx 80$ years. Precession period in this binary, caused by the gravitational perturbation of the contact binary, is estimated to be about 120 years. The wide mutual orbit and orbit of the non-eclipsing pair are found to be close to coplanarity, preventing any changes of the inclination angle of the non-eclipsing orbit and excluding occurrence of the second system of eclipses in future. Possibilities of astrometric solution and direct resolving of the wide, mutual orbit are discussed. Nearby star, HD95606, was found to form loose binary with quadruple system VW~LMi. ", "introduction": "\\label{intro} The photometric variability of VW~LMi (HIP~54003, HD~95660, sp. type F3-5V, $V_{max}$=8.0), was found by the Hipparcos mission \\citep{hipp}, where it was correctly classified as a W~UMa-type eclipsing binary with an orbital period of 0.477547 days. The first ground-based photometric observations of the system obtained in 1999 and 2000 (taken in the $B$ and $V$ Johnson filters) were published by \\citet{dumi2000}. Analysis of these light curves \\citep{dumi2003} lead to the photometric mass ratio $q_{ph}$ = 0.395 and inclination $i$ = 72.4\\degr and contact configuration for the system. Later \\citet{gome2003} presented new $BV$ photometry and its preliminary analysis. Assuming convective envelopes for both components and the temperature of the primary as $T_1$ = 6700 K the authors estimated the mass ratio as $q=0.4$, and inclination around 70\\degr. Fourier analysis of its Hipparcos light curve (hereafter LC) presented by \\citet{sela2004} yielded quite different parameters: $q = 0.25$, $i$ = 72.5\\degr~and fill-out factor $f=0.4$. The discovery of the second (non-eclipsing) binary in VW~LMi by \\citet{ddo11} makes all previous photometric solutions almost useless due to strong light contribution of the second pair of about $(L_3 + L_4)/(L_1 + L_2) = 0.42$ (at the maximum brightness of the contact pair) which was not taken into account. \\citet{ddo11} presented long-term spectroscopy (209 spectra taken between 1998 and 2005) of the system obtained at David Dunlap Observatory (hereafter DDO) which enabled to disentangle all three orbits in this tight multiple system: the contact binary with the $P_{12}$ = 0.4775 days period is orbiting another binary with $P_{34} = 7.93$ days in a relatively tight, 355-days, mutual orbit. Using preliminary inclination angle of the contact-binary orbit found by photometric analysis, $i_{12}$ = 80.1$\\degr$, the authors determined masses of all components and found that the orbits of the binaries are not coplanar. Light-time effect (hereafter LITE) with peak-to-peak range of $2A$ = 0.0074 days was predicted to be seen in the minima of the contact pair as a result of the mutual revolution of the binaries. The LITE was found in published minima by \\citet{bajo2007}. The corresponding orbital parameters, $A$ = 0.0037(4) days, $P_{1234}$ = 353(2) days, $e$ = 0.5(2) and $\\omega$ = 2.8(4) rad, are rather preliminary due to few available minima. The eccentricity corresponding to their orbital solution is much higher than predicted spectroscopically by \\citet{ddo11}. The spectral type of VW~LMi was estimated as F5V by \\citet{ddo11}, while observed Tycho-2 $(B-V)$ = 0.21 and 2MASS $J-K$ = 0.34 colors correspond to F2V spectral type. Both determined spectral type and colors refer to the whole quadruple system. VW~LMi is tightest quadruple system known \\citep{toko2008}. Also it has the shortest period of the outer orbit within multiple systems harboring contact binaries. The ratio of the outer orbital period and orbital period of the non-eclipsing pair is only about $P_{1234}/P_{34}$ = 44.5 hence we can expect secular orbital changes on the timescales as short as decades. The chances of resolving of components (binaries) by either speckle or long-baseline interferometry are rather meager: the maximum angular separation of the components was estimated to be only about 10mas \\citep{ddo11}. Astrometric observations of VW~LMi do not indicate its multiplicity. Goals of the present paper are as follows: (i) to present and analyze new photometric and spectroscopic observations, (ii) to perform simultaneous solution of LITE using both minima times of the contact binary and radial velocities (hereafter RVs) of the individual components, (iii) to assess possibility of the tidal disturbances of the inner orbits and resulting precession, (iv) determine absolute parameters of all four components. ", "conclusions": "Tightest known quadruple system VW~LMi is really unique. It definitely deserves additional observations and analysis. The study of this system could bring light on (i) the evolution and origin of binary stars in multiple systems of stars (ii) tidal interaction of third body and its influence on the Roche geometry (iii) long-term evolution of orbits in tight multiple systems. The system is useful for further analysis since all four components are visible in the spectrum. The determination of the individual component's parameters like $T_{eff}$, $\\log g$, metalicity could benefit from spectra disentangling (see \\citet{hadr1995}). The absolute parameters of all components which had to originate at practically same time could be used to determine the evolutionary status and history of the system. The study of VW~LMi is, however, complicated by (i) the outer orbital period being close to one year which complicates its phase coverage by Earth-bound observer (ii) very small angular separation of the components making direct resolving of the mutual wide orbit and reliable determination of $i_{1234}$ difficult (which would definitely improve determination of individual masses), (iii) fast rotation of the components of the contact binary making RV measurements unsure and rectification to real continuum impossible (due to the line-blanketing) (iv) eclipses and Roche geometry in the contact pair making usual assumption of the spectra disentangling techniques not valid (line profiles of individual components cannot change with orbital revolution). It is also interesting to note, that VW~LMi and HD95606 show very similar proper motion, parallax and RVs (see Table~\\ref{tab02}) - the stars definitely form a loosely bound pair and all components very probably evolved from the same protostellar cloud (see \\citet{oswa2007}). The orbit of the detached pair in VW~LMi with 7.93-days period is almost circular. This is the case of another two quadruples detected by the DDO observations, TZ~Boo and V2610~Oph (DDO series No. XIV, \\citet{ddo14}). That means that either it evolved with such orbit or it was circularized by the gravitational interaction with the contact binary. Detected apsidal motion in VW~LMi requires further observing to reliably determine apsidal period. Investigation of the observed eccentricities of the second binaries and their predicted synchronization timescales in quadruple systems with contact binaries could shed light on the age and evolutionary status of contact binaries \\citep{ruci-priv2008}. Better characterization of VW~LMi calls for long-term monitoring to cover the whole 355-days orbital cycle. Especially, times of minima should be obtained free of any systematic effects. The understanding of the system would greatly benefit from visual orbit obtained by means of long-baseline interferometry. Multi-color photometry and/or echelle spectroscopy could lead to reliable determination of component's temperatures and luminosities. \\medskip The stays of TP at DDO have been supported by a grant to Slavek M. Rucinski from the Natural Sciences and Engineering Council of Canada. This research has been supported in part by the Slovak Academy of Sciences under grants No. 2/7010/7 and 2/7011/7, and grant of the \\v{S}af\\'arik University VVGS 9/07-08. MV's research is supported by a Marie Curie ``Transfer of Knowledge'' Fellowship within the 6th European Community Framework Programme. The observations at Astronomical Observatory at Kolonica Saddle (part of Vihorlat Observatory) were partially supported by APVV grant LPP-0049-06 and APVV bilateral grant SK-UK-01006. DB thanks to Miron Kerul-Kmec for technical assistance with CCD camera at the Roztoky Observatory." }, "0808/0808.0403_arXiv.txt": { "abstract": "{The magnetorotational instability (MRI) of differential rotation under the simultaneous presence of axial and azimuthal components of the (current-free) magnetic field is considered. For rotation with uniform specific angular momentum the MHD equations for axisymmetric perturbations are solved in a local short-wave approximation. All the solutions are overstable for $B_z \\cdot B_\\phi\\neq 0$ with eigenfrequencies approaching the viscous frequency. For more flat rotation laws the results of the local approximation do not comply with the results of a global calculation of the MHD instability of Taylor-Couette flows between rotating cylinders. -- With $B_\\phi$ and $B_z$ of the same order the traveling-mode solutions are also prefered for flat rotation laws such as the quasi-Kepler rotation. For magnetic Prandtl number ${\\rm Pm}\\to 0$ they scale with the Reynolds number of rotation rather than with the magnetic Reynolds number (as for standard MRI) so that they can easily be realized in MHD laboratory experiments. -- Regarding the nonaxisymmetric modes one finds a remarkable influence of the ratio $B_\\phi/B_z$ only for the extrema. For $B_\\phi\\gg B_z$ and for not too small Pm the nonaxisymmetric modes dominate the traveling axisymmetric modes. For standard MRI with $B_z\\gg B_\\phi$, however, the critical Reynolds numbers of the nonaxisymmetric modes exceed the values for the axisymmetric modes by many orders so that they are never prefered. } ", "introduction": "It has been shown in previous publications starting with Hollerbach \\& R\\\"udiger (2005) that the magnetorotational instability (MRI) under the presence of both current-free axial and azimuthal components of the magnetic field ('Helical' fields, hence HMRI) is always characterized by an eigenoscillation frequency. In combination with the vertical wavenumber the resulting instability pattern is thus an axisymmetric wave traveling along the rotation axis. In all our considerations the magnetic Prandtl number Pm plays the basic role. For $\\rm Pm \\to 0$ the HMRI scales with the Reynolds number Re rather than with the magnetic Reynolds number Rm as it does for the standard magnetorotational instability for an axial magnetic field. The questions arise whether this frequency reflects the geometry of the magnetic field and whether it is observable with real astrophysical objects such as protoneutron stars and/or accretion disks. Also the relation to the Azimuthal MagnetoRotational Instability (AMRI, see R\\\"udiger et al. 2007b) for (current-free) toroidal fields which is nonaxisymmetric and which scales with the magnetic Reynolds number for $\\rm Pm \\to 0$ must be considered. We find the AMRI only weakly (if ever) influenced by the addition of an axial field which is not much stronger than the toroidal field. For $\\rm Pm\\simeq 1$ the difference between Re and Rm disappears so that the main differences of the instabilities also disappear. In the present paper we start with a local approximation using analytical methods for the most simple rotation law for constant specific angular momentum, i.e. the Rayleigh limit. Global calculations of the stability of the same rotation law between two rotating perfect-conducting cylinders and threaded by a helical current-free magnetic field lead to an almost perfect coincidence of the results of both methods. This is no longer true, however, for more flat rotation laws such as quasikeplerian rotation in a finite gap where the differences of the short-wave approximation (which only holds for infinitely thin gaps) and global models are so strong that the results completely differ (R\\\"udiger \\& Hollerbach 2007). ", "conclusions": "" }, "0808/0808.0918_arXiv.txt": { "abstract": "We present stellar velocity dispersion measurements in the host galaxies of 10 luminous quasars (M$_{V}<-$23) using the Ca H\\&K lines in off-nuclear spectra. We combine these data with effective radii and magnitudes from the literature to place the host galaxies on the Fundamental Plane (FP) where their properties are compared to other types of galaxies. We find that the radio-loud (RL) QSO hosts have similar properties to massive elliptical galaxies, while the radio-quiet (RQ) hosts are more similar to intermediate mass galaxies. The RL hosts lie at the upper extreme of the FP due to their large velocity dispersions ($\\langle \\sigma_{*} \\rangle$ = 321~km~s$^{-1}$), low surface brightness ($\\langle \\mu_{e}(r) \\rangle$ = 20.8~mag~arcsec$^{-2}$), and large effective radii ($\\langle R_{e} \\rangle$ = 11.4~kpc), and have $\\langle M_{*} \\rangle$ = 1.5 x 10$^{12}$ M$_{\\sun}$ and $\\langle M/L \\rangle$ = 12.4. In contrast, properties of the RQ hosts are $\\langle \\sigma_{*} \\rangle$ = 241~km~s$^{-1}$, $\\langle M_{*} \\rangle \\sim$ 4.4 x 10$^{11}$ M$_{\\sun}$, and M/L $\\sim$ 5.3. The distinction between these galaxies occurs at $\\sigma_{*}\\sim$~300~km~s$^{-1}$, R$_{e} \\sim$ 6~kpc, and corresponding M$_{*} \\sim$ 5.9 $\\pm$ 3.5 x 10$^{11}$ M$_{\\sun}$. Our data support previous results that PG QSOs are related to gas-rich galaxy mergers that form intermediate-mass galaxies, while RL QSOs reside in massive early-type galaxies, most of which also show signs of recent mergers or interactions. Most previous work has drawn these conclusions by using estimates of the black hole mass and inferring host galaxy properties from that, while here we have relied purely on directly measured host galaxy properties. ", "introduction": "A growing understanding of the connection between galaxies and their central black holes has emerged over the last decade. We now know that all galaxies with a bulge contain supermassive black holes \\citep{kormendy04} and that black hole mass is correlated with host galaxy stellar velocity dispersion \\citep{gebhardt00a,ferrarese00,tremaine02}. Furthermore, the inclusion of an amount of energy equal to that expected from AGN feedback to quench star formation above a critical halo mass in semi-analytic galaxy formation models \\citep{cattaneo06,dekel06} reproduces the galaxy demographics and bimodality of properties observed in large surveys (SDSS: \\citet{kauffmann03a,kauffmann03b,hogg03,baldry04,heavens04,cidfernandes05}; GOODS: \\citet{giavalisco04}; COMBO-17: \\citet{bell04}; DEEP/DEEP2: \\citet{koo03,koo05,faber07}; MUNICS: \\citet{drory01}; FIRES: \\citet{labbe03}; K20: \\citet{cimatti02}; GDDS: \\citet{mccarthy04}). These facts suggest that the growth mechanisms of the black hole and galaxy must be connected. However, details of the physical processes that make this connection, such as how AGN energy interacts with and is dissipated by surrounding halo gas, are not yet understood. One way to investigate these processes is to understand the nature of the host galaxies. Does something in the galaxy trigger AGN activity? Do active quasars exist in galaxies with similar properties? Studies of AGN host galaxies have reached different conclusions. One group of collaborators, \\citep{MKDBOH99,HKDB00,NDKHBJ00} believes these objects to be predominantly normal massive ellipticals, including Nolan et al. (2001) who found that most quasar host galaxies had evolved stellar populations, 10~Gyr old, with only a very small amount of recent star formation. However, \\citet{mil81}, in the first spectroscopic investigation of a sample of these objects, concluded that they are not normal luminous ellipticals, a result which was later confirmed with deeper spectroscopy from the Keck telescope \\citep{mts96,she01,miller03}. Moreover, \\citet{CS00,CS01} have seen evidence of star formation within the past 100~Myr in quasars hosts with far-infrared excesses using deep spectra from Keck. There is still debate about whether the different results are due to better quality spectra taken closer to the nucleus using 8-10~m class telescopes, or whether real differences exist in the host galaxy properties of the different quasar samples studied \\citep{lacy06}. This question will no doubt be answered as more data are analyzed from the larger extra-galactic surveys. In this work we study luminous quasars (M$_{V}<-$23) in which the galaxy is actively feeding the central supermassive black hole. It is here that we should be able to investigate connections between the black hole and its surrounding galaxy from which we can draw conclusions about how the black hole may or may not affect galaxy formation and evolution. We begin in this first paper by analyzing the structural properties of the QSO host galaxies with the use of directly measured stellar velocity dispersions, previously unobtainable for quasars this luminous. We use these data to place the host galaxies on the Fundamental Plane and ascertain their structural properties relative to other types of galaxies. In future papers we will investigate whether these objects follow the M$_{BH}$-$\\sigma$ relation and analyze the host galaxy stellar populations to look for indications of star formation activity relative to quasar activity. This paper is organized in the following manner. In \\S \\ref{data_section} we describe the sample selection and data analysis, including our removal of scattered quasar light from the observed spectra and the measurement of stellar velocity dispersions. We also present the comparison objects from the literature that are used in our analysis. In \\S \\ref{result_section} we use the fundamental parameters derived in \\S \\ref{data_section} to analyze the Fundamental Plane locations and mass-to-light ratios of these objects. In \\S \\ref{discussion} we discuss the properties of our QSO host galaxies relative to the comparison objects and their implication that two different classes of objects are present in our sample. Finally, \\S \\ref{summary} summarizes our work and presents the main conclusions. Further details about our stellar velocity dispersion measurement limitations and potential biases can be found in Appendix \\ref{bias}. ", "conclusions": "} We have for the first time directly measured host galaxy stellar velocity dispersions for very luminous (M$_{V}<-$23) quasars, including both radio-loud and radio-quiet objects, and analyzed their structural properties. We compare the properties of these host galaxies to those of normal early-type galaxies \\citep{bernardi03a}, giant early-type galaxies \\citep{bernardi06}, a sample of radio-quiet PG QSO hosts \\citep{dasyra07}, and galaxy merger remnants \\citep{rothberg06}. The following summarizes our main conclusions. \\begin{enumerate} \\item{ The six radio-loud QSO host galaxies lie at the upper extreme the Fundamental Plane of early-type galaxies. They occupy this location due to their large velocity dispersions (with an average of 321~km~s$^{-1}$), large effective radii (average of 11.4 kpc), low surface brightness (average of 20.8~mag~arcsec$^{-2}$), and high M/L (average of 12.4). The properties of these radio-loud host galaxies are similar to those of giant early-type galaxies in the SDSS, although only one has the spectrum of a purely old giant elliptical galaxy. } \\item{ The four radio-quiet QSO host galaxies reside on the Fundamental Plane among normal early-type galaxies and at the high end of other PG QSO hosts, with a mean velocity dispersion of 241~km~s$^{-1}$. Their surface brightness and effective radii are slightly higher than normal early-type galaxies. The M/L's are slightly below normal early-type galaxies. } \\item{ The radio-loud hosts in our study are either elliptical galaxies or interacting, while the radio-quiet hosts show a mixture of spiral and elliptical structure. The distinction between the two groups is due to galaxy mass, inferred from measured structural parameters, rather than morphological galaxy type. The separation occurs at galaxy velocity dispersions of $\\sigma_{*}\\sim$~300~km~s$^{-1}$, effective radii of R$_{e} \\sim$ 6~kpc, and corresponding stellar masses of M$_{*} \\sim$ 5.9 $\\pm$ 3.5 x 10$^{11}$~M$_{\\sun}$. } \\item{ We confirm a correlation between radio luminosity and stellar velocity dispersion, and thus implied black hole mass, of the host galaxies that suggests a higher slope (L$_{5 GHz}\\varpropto$~M$_{BH}^{3.8}$ with rms scatter of 1.36 dex) than found by Franceschini et al.~(L$_{5 GHz}\\varpropto$~M$_{BH}^{2.66}$), though it could be consistent with previous work given the large scatter in this relation \\citep{mclure04}, our small sample size, the limited M$_{BH}$ range of our data, and differences in M$_{BH}$ estimated from different techniques. We find a tighter correlation between radio luminosity and host galaxy bulge mass, $L_{radio} \\sim M_{bulge}^{3.56}$ with rms scatter of 1.09 dex } \\end{enumerate}" }, "0808/0808.2630_arXiv.txt": { "abstract": "{} {We seek to determine whether the late-type star 2MASS J12354893$-$3950245 (2M1235$-$39) is a member of the TW Hya Association (TWA), a hypothesis suggested by its association with a bright X-ray source detected serendipitously by ROSAT and XMM-Newton and its ($\\sim3'$) proximity to the well-studied (A+M binary) system HR 4796.} {We used optical spectroscopy to establish the Li and H$\\alpha$ line strengths of 2M1235$-$39, and determined its proper motion via optical imaging. We also considered its X-ray and near-IR fluxes relative to the M star HR 4796B.} {The optical spectrum of 2M1235$-$39 displays strong Li absorption and H$\\alpha$ emission (equivalent widths of 630 m\\AA\\ and $-6.7$ \\AA, respectively). Comparison of the spectrum with that of a nearby field star, along with the DENIS catalog $IJK$ magnitudes, indicates the spectral type of 2M1235$-$39 is M4.5. We measure a proper motion for 2M1235$-$39 that agrees, within the errors, with that of HR 4796.} {The Li absorption and H$\\alpha$ emission line strengths of 2M1235$-$39, its near-IR and X-ray fluxes, and its proper motion all indicate that 2M1235$-$39 is a TWA member. Most likely this star is a wide (13,500 AU) separation, low-mass, tertiary component of the HR 4796 system.} ", "introduction": "As of little more than a decade ago, astronomers were almost oblivious to the presence of low-mass, pre-main sequence stars within $\\sim100$ pc of Earth. The intervening years have seen the identification of a few hundred such stars, with ages ranging from 8 to 100 Myr, as part of numerous post-T Tauri associations (Zuckerman \\& Song 2004, hereafter ZS04, and references therein; Torres et al.\\ 2006, 2008). Perhaps the greatest excitement associated with the recognition of the existence of nearby young stars has been the opportunity to study, at close range, the evolution of youthful planetary systems, via direct thermal imaging of warm massive planets (e.g., Chauvin et al.\\ 2004; Song et al.\\ 2006) and via imaging and spectroscopy of debris disks (e.g., Rebull et al.\\ 2008 and references therein). Young, local stellar groups also afford unique insight into the early evolution of low-mass stars and ultracool dwarfs (e.g., Looper et al.\\ 2007; Cruz et al.\\ 2008; and references therein). The difficulty inherent in identifying young stars and young star groups near Earth reflects the fact that such groups are spread over large areas of the sky (ZS04). Furthermore, while the local young groups are usually ``spearheaded'' by a handful of well-studied, individual systems that feature, e.g., strong H$\\alpha$ emission, enormous IR excesses, and/or easily imaged debris disks (TW Hya and $\\beta$ Pic being cases in point), the vast majority of nearby young stars are otherwise unremarkable late-type (K through M) dwarfs that do not stand out or even turn up in optical emission-line or far-infrared (e.g., IRAS) surveys. However, all $\\sim10$--100 Myr-old stars of types F through M are at or near the peaks of their lives in terms of their X-ray luminosities relative to bolometric (with ``saturated'' values of $L_X/L_{\\rm bol} \\sim 10^{-3}$, ZS04 [their Fig.~4]; see also Kastner et al.\\ 1997 and Preibisch \\& Feigelson 2005). Hence, X-ray point source catalogs, in tandem with recently released, comprehensive catalogs of distances and proper motions of stars in the solar neighborhood, have served as the main resources with which to isolate stars that are likely nearby and young. Followup optical spectroscopy and/or imaging then readily confirms (or refutes) membership in the ``nearby young star club,'' via determination of surface Li abundances and relative ($UVW$) Galactic space motions. Here, we demonstrate that serendipitous XMM-Newton and ROSAT X-ray detections of 2MASS J12354893$-$3950245 (hereafter 2M1235$-$39), combined with its optical spectrum and proper motion, establishes this star as a member of the quintessential local young star group, the TW Hya Association (TWA; Kastner et al.\\ 1997; Webb et al.\\ 1999; Zuckerman et al.\\ 2001). Indeed, 2M1235$-$39 is, likely, the tertiary component of the well-studied HR 4796 (A+M star) binary system (Jura et al.\\ 1993; Stauffer et al.\\ 1995), which is designated TWA 11. ", "conclusions": "All measured properties of 2M1235$-$39 presented here --- its Li absorption and H$\\alpha$ emission line strengths, its near-IR and X-ray fluxes, and its proper motion --- are compatible with TWA membership. Based on these results and on the similarity of its common proper motion to that of HR 4796A, we conclude that 2M1235$-$39 is most likely a wide (13,500 AU) separation, low-mass, tertiary component of the HR 4796 system." }, "0808/0808.0773_arXiv.txt": { "abstract": "We show that the smoothed particle hydrodynamics (SPH) method, used with individual time-steps in the way described in the literature, cannot handle strong explosion problems correctly. In the individual time-step scheme, particles determine their time-steps essentially from a local Courant condition. Thus they cannot respond to a strong shock, if the pre-shock timescale is too long compared to the shock timescale. This problem is not severe in SPH simulations of galaxy formation with a temperature cutoff in the cooling function at $10^4~{\\rm K}$, while it is very dangerous for simulations in which the multiphase nature of the interstellar medium under $10^4~{\\rm K}$ is taken into account. A solution for this problem is to introduce a time-step limiter which reduces the time-step of a particle if it is too long compared to the time-steps of its neighbor particles. Thus this kind of time-step constraint is essential for the correct treatment of explosions in high-resolution SPH simulations with individual time-steps. ", "introduction": "Hierarchical (individual) time-step method \\citep[e.g.,][]{McMillan1986, HernquistKatz1989, Makino1991IndividualTimeStep} is widely used in simulations of galaxy formation and star formation based on smoothed particle hydrodynamics (SPH) method \\citep{Lucy1977, GingoldMonaghan1977}. This method allows particles to have different time-steps, and can significantly reduce the total calculation cost when there is a large variation in the timescales of particles. Almost all implementations of the individual time-steps method used for particle systems violates Newton's third law (see \\citeauthor{FarrBerschinger2007} \\citeyear{FarrBerschinger2007} for one of exception). As long as physical quantities are integrated with sufficient accuracy, this violation is not a severe problem. However, it is not always possible to maintain the accuracy. To our knowledge, all existing implementations of individual time-step for SPH rely on the determination of the time-step at the end of previous time-steps. Therefore, if something unforeseen occurs during the time-step of one particle, the particle might fail to catch that event, resulting in a large integration error. A supernova (SN) explosion is an example of such an event. A SN generates a small amount of very hot gas ($T \\sim 10^8~{\\rm K}$) in a large clump of cold gas ($T \\sim 10~{\\rm K}$). The difference in the time-steps of the hot gas and surrounding cold gas particles becomes quite large (typically the difference reaches $\\sim 10^3$). Thus, hot gas particles step forward $\\sim 10^3$ or more time-steps before neighboring cold gas particles respond to the SN event. This means that the evolution of both hot and cold gas particles is completely wrong, since the surrounding cold gas particles do not react the explosion for a duration much longer than the timescale in which the blast wave would propagate the inter-particle distance. This problem is not severe for ordinary SPH simulations of galaxy formation because of the temperature cutoff in cooling functions at $10^4~{\\rm K}$, while it becomes very serious for simulations involving the multiphase nature of the interstellar medium under $10^4~{\\rm K}$, because the mach number can be very high. This problem of sudden change occurs in any dynamical simulation with individual time-steps, as long as the time-steps are determined with the usual explicit method. In principle, a fully implicit method in which the time-step itself is also determined implicitly \\citep{Makino+2006} or a method which satisfies Newton's third law \\citep{FarrBerschinger2007} can solve this problem, but there are no implementations of such methods for SPH simulations yet {\\footnote {Recently, \\cite{Springel2009} has developed a mesh-based scheme employing individual time-steps where the smaller time-step is adopted for integrations between neighboring meshes. This scheme does satisfy Newton's third law.}}. In simulation of star clusters, the SN and its kick introduces a sudden change in the orbit (and mass) of the exploded star. Here, a rather simple prescription in which either all stars or at least nearby stars are synchronized to the time of explosion and restart the integration has been used. This prescription, at least the version which synchronizes all particles in the system, is impractical for $N$-body/SPH simulations of galaxies, because the number of SN events and therefore the increase in the calculation cost is too great. We propose a simple limiter for hydrodynamical time-steps in the individual time-step method which mitigates this problem. With this limiter the behavior of an explosion integrated by individual time-steps becomes essentially the same as that integrated by global time-steps. In \\S 2, we describe this limiter, and in \\S 3 we report the result of numerical experiments. ", "conclusions": "" }, "0808/0808.1668_arXiv.txt": { "abstract": "We carried out a comparison of the signals seen in contemporaneous BiSON and GOLF data sets. Both instruments perform Doppler shift velocity measurements in integrated sunlight, although BiSON perform measurements from the two wings of potassium absorption line and GOLF from one wing of the NaD1 line. Discrepancies between the two datasets have been observed. We show,in fact, that the relative power depends on the wing in which GOLF data observes. During the blue wing period, the relative power is much higher than in BiSON datasets, while a good agreement has been observed during the red period. ", "introduction": "P-modes are standing acoustic waves in the solar interior. There are instances when it has been shown that solar flares are correlated with modes of oscillation in the Sun just as the Earth is set ringing after a major earthquake. These events have been seen in high-degree modes (\\cite{fog},\\cite{kos}). We are interested in the possibility that the global modes are similarly stimulated and we seek to use the extensive BiSON dataset to search for these events. Large power events with strengths many times the mean level are seen and, moreover, the number is in excess of the predictions of stochastic excitation \\cite{Cha1}. Our long-term aim is to categorize these rare events through their temporal features in both power and phase with the hope of being able to distinguish natural from forced events. As part of this process of categorization of large excitations, we carried out a comparison of the signals seen in contemporaneous BiSON and GOLF data. The intention is to show that the power and the phase returned for the signal is in agreement for the two datasets. Given that, we can go on to use the very extensive BiSON data on its own. We show here that relative power in the two datasets depends on the wing in which the GOLF data are observed. We compare the observed power ratio to the predictions based on the heights in the atmosphere that the respective spectral lines are formed. We show also that the relative phases returned by the analysis are comparable. ", "conclusions": "We have used the comparison of contemporaneous GOLF and BiSON data to demonstrate that the power and phase evolution of the solar oscillations data are consistent between the two datasets and are therefore likely to be true characteristics of the oscillations of the Sun and not instrumental in origin. The red-wing GOLF data were found to be most useful for this. The BiSON data were shown not to be compromised by their lack of perfect fill nor their ground-based observations. This opens up the possibility to use the full BiSON dataset spanning three solar cycles in duration for further study of the large excitations and their links with solar activity. The next step in this project is to continue with the categorization of events with the eventual hope of being able to identify those that are stimulated by activity on the Sun." }, "0808/0808.3806_arXiv.txt": { "abstract": "{ In an attempt of clarifying the connection between the photospheric abundance anomalies and the stellar rotation as well as of exploring the nature of ``normal A'' stars, the abundances of seven elements (C, O, Si, Ca, Ti, Fe, and Ba) and the projected rotational velocity for 46 A-type field stars were determined by applying the spectrum-fitting method to the high-dispersion spectral data obtained with BOES at BOAO. We found that the peculiarities (underabundances of C, O, and Ca; an overabundance of Ba) seen in slow rotators efficiently decrease with an increase of rotation, which almost disappear at $v_{\\rm e}\\sin i \\ga 100$~km~s$^{-1}$. This further suggests that stars with sufficiently large rotational velocity may retain the original composition at the surface without being altered. Considering the subsolar tendency (by several tenths dex below) exhibited by the elemental abundances of such rapidly-rotating (supposedly normal) A stars, we suspect that the gas metallicity may have decreased since our Sun was born, contrary to the common picture of galactic chemical evolution. } ", "introduction": "Since unevolved A-type stars on (or near to) the upper main-sequence have masses around $\\sim 2 M_{\\odot}$, their surface abundances may retain information of the composition of the past galactic gas ($\\la 10^{9}$~yr ago) from which they formed; this would provide us with an important opportunity to investigate the ``recent'' chemical evolution history of the Galaxy. However, such a study using A stars as a probe of late-time history of chemical evolution has rarely been done in spite of its potential significance\\footnote{For example, Takeda, Sato, \\& Murata (2008) found in their extensive study of late-G giants (also having masses around $\\sim 2 M_{\\odot}$; evolved counterparts of A dwarfs) that their metallicities show an appreciable diversity as large as $\\sim 1$~dex ($-0.8 \\la$~[Fe/H]~$\\la +0.2$) with a subsolar trend on the average (cf. Fig. 14 therein), from which they argued that some special event (like a mixing of metal-poor primordial gas caused by infall) might have occurred $\\sim 10^{9}$~yr ago. However, before making any speculation, it is important to confirm whether such a trend is also observed in their progenitors in the upper main-sequence.} in contrast to the case of old-time history where a number of researches (using longer-lived F--G--K dwarfs) are available. This is related to the fact that a large fraction of them are rapid rotators (typically $v_{\\rm e} \\sin i \\sim$ 100--200~km~s$^{-1}$ on the average; see, e.g., Royer et al. 2002a, b) whose spectra are technically difficult to analyze because lines are broad and smeared out, while sharp-lined slow rotators easy to handle tend to show abundance anomalies (chemically peculiar stars or CP stars).\\footnote{Although many things are left unresolved concerning the origin and nature of the CP phenomena, it is widely considered (at least in the qualitative sense) that the chemical segregation in the stable atmosphere is responsible for the abundance anomalies at the surface: i.e., an element becomes over- or under-abundant depending on the balance of upward radiation force and the downward gravitational force. In this case, rotation would act against an efficient built-up of such anomalies because it enhances mixing of outer stellar layers via shear instability or meridional circulation.} As a matter of fact, most spectroscopic analyses of A stars have focused on sharp-lined ones with $v_{\\rm e}\\sin i \\la 50$~km~s$^{-1}$. For this reason, nobody could be sure whether the result obtained for star classified as ``normal A'' really reflects the initial composition free from any peculiarities. Therefore, in order to make a further step forward, it is requisite to challenge abundance determinations for ``unbiased'' sample of A-type stars in general (i.e., without sidestepping rapidly rotating ones), which inevitably requires an application of the spectrum synthesis technique, since reliably measuring the equivalent widths of individual spectral lines is almost hopeless for rapid rotators. Admittedly, while such trials of determining abundances from spectra of A dwarfs including broad-lined ones have recently emerged thanks to the improvement in the method of analysis as well as the data quality, their interests are mainly directed to objects of specific types; e.g., Vega-like stars (Dunkin et al. 1997) or $\\lambda$ Bootis stars (Andrievsky et al. 2002) or open-cluster stars (Takeda \\& Sadakane 1997; Varenne \\& Monier 1999; Gebran et al. 2008; Gebran \\& Monier 2008; Fossati et al. 2008). Namely, a systematic study attempting to clarify the characteristics of normal field A-type stars in general, especially in terms of their abundance--rotation connection, seems to have been rarely attempted. To our knowledge, only one such study is Lemke's (1990, 1993) determinations of C and Ba abundances for some 20 rapidly-rotating A stars with $v_{\\rm e} \\sin i$ up to $\\sim$~200~km~s$^{-1}$, which however appear to be still insufficient and inconclusive as judged from his adopted method of approach as well as the number of elements studied. Considering this situation, we decided to revisit this problem in our own manner based on the high-dispersion spectral data of $\\sim 50$ A-type stars in a wide range of $v_{\\rm e} \\sin i$ (0--300~km~s$^{-1}$) obtained with BOES at BOAO, while applying the automatic spectrum fitting algorithm (Takeda 1995) which efficiently enables determinations of the abundances (for selected six elements of C, O, Si, Ti, Fe, and Ba) even for rapid rotators showing considerably merged spectra. Our ultimate aim is to clarify the following questions of interest:\\\\ --- (1) Is there any systematic rotation-dependent tendency between slow and rapid rotators in terms of the abundance anomaly? If so, what is the critical value of $v_{\\rm e} \\sin i$, above which stars may be regarded as normal? \\\\ --- (2) What would the abundance characteristics of ``normal A-type stars'' like, which we may consider as retaining the composition of the galactic gas from which they formed? \\\\ We will show that reasonable answers to these points are provided from this study (cf. Sect. V). ", "conclusions": "\\subsection{Rotation--Abundance Connection} Figures 9a--f display the resulting [X/H] vs. [Fe/H] correlations (for X = C, O, Si, Ca, Ti, and Ba), from which we can roughly divide these elements into three groups.\\\\ (i) Si and Ti: almost scaling in accordance with Fe.\\\\ (ii) C, O, and Ca: showing an anti-correlation trend with Fe.\\\\ (iii) Ba: positive correlation with Fe, though its range of peculiarity is much more conspicuous than that of Fe. Besides, [C/H], [O/H], [Si/H], [Ca/H], [Ti/H], [Ba/H], and [Fe/H] are plotted against $v_{\\rm e} \\sin i$ in Figures 10a--g. We can recognize from these figures that [C/H], [O/H], [Ca/H] and [Ba/H] are systematically $v_{\\rm e} \\sin i$-dependent in the sense that the peculiarity (overabundance for Ba, underabundance for C/O/Ca) tends to decrease with an increase in $v_{\\rm e} \\sin i$. While such a convincing tendency is not apparent for the remaining elements (Si, Ti, and Fe), [Fe/H] appears to weakly conform to this trend (i.e., decreasing tendency with $v_{\\rm e} \\sin i$). Combining these observational fact, we may conclude as follows:\\\\ --- (a) All the seven elements exhibit some kind of abundance peculiarities, which are more conspicuously seen in slow rotators ($v_{\\rm e}\\sin i \\la 50$~km~s$^{-1}$) and characterized by the deficiency of C, O, and Ca and the enrichment of Si, Fe, and (especially) Ba.\\\\ --- (b) These anomalies tend to diminish progressively with an increase in $v_{\\rm e} \\sin i$ (at least in the range of slow/moderate rotators of $\\la 100$~km~s$^{-1}$). \\\\ --- (c) The stellar rotational velocity must thus be the most important key factor in the sense that the extent of abundance peculiarity tends to be larger as a star rotates more slowly, which is presumably because some counter-acting mechanism of diluting the built-up anomaly (most probably due to the element segregation in a stable atmosphere/envelope) takes place in rapid rotators. We also point out these tendencies seen in Figures 9 and 10 are more or less consistent with the results of recently published papers focused on the abundance trends of A-type stars (including Am stars) for a wide range of $v_{\\rm e} \\sin i$ values: e.g., Lemke (1990, 1993) [field stars; C, Ba (elements in common with this study)], Savanov (1995a,b) [field stars; C, O, Si, Ca, Fe, Ba], Takeda \\& Sadakane (1997) [Hyades and field stars; Fe, O], Gebran, Monier, \\& Richard (2008) [Coma Berenices; C, O, Si, Ca, Fe, Ba], Gebran \\& Monier (2008) [Pleiades; C, O, Si, Ca, Fe, Ba], and Fossati et al. (2008)[Praesepe; C, O, Si, Ca, Fe, Ba]. \\subsection{Implication of Subsolar Compositions in Normal A Stars} According to what we learned in Sect. V-a, we may assume that the abundance peculiarities of A-type stars (conspicuously seen slow rotators) tend to disappear for rapid rotators at $v_{\\rm e} \\sin i \\ga 100$~km~s$^{-1}$ (cf. Figure 10). If so, we would be able to gain information of the galactic gas $\\la 10^{9}$~yr ago by inspecting the photospheric abundances of such rapidly-rotating A-type stars, since they are considered to retain the composition of the gas from which they formed. From this point of view, it is interesting to note in Figure 10 that the [X/H] values at the high-$v_{\\rm e} \\sin i$ range tend to be somewhat negative or ``subsolar'' for many elements such as C, O, Ti, Fe, and Ba; i.e., by several tenths dex below the solar (or Procyon) abundances on the average. Here we recall Takeda, Sato, \\& Murata's (2008) conclusion that the [Fe/H] values (as well as those of other elements whose abundances almost scale with Fe) of evolved G giants, many of which have mass values around $\\sim 2 M_{\\odot}$ like A-type dwarfs, spread in a range of $-0.8 \\la$~[Fe/H]~$\\la +0.2$ around an average value of [Fe/H]~$\\sim -0.3$. Considering these two observational consequences, we would conclude that the metallicities of the galactic gas $\\la 10^{9}$~yr ago had really a subsolar tendency (though with a rather large diversity). If this is the case, the gas metallicity of [Fe/H] $\\sim 0$ ($\\sim 5 \\times 10^{9}$ ago when our Sun was born) must have decreased by several tenths dex with an elapse of time until $\\la 10^{9}$~yr ago when A dwarfs (progenitors of G giants) were born. Although this trend does not seem to have been taken very seriously so far \\footnote{Meanwhile, a completely different solution to this problem has also been proposed, arguing the necessity of downward revision of the solar abundances as a result of the application of sophisticated 3D line formation theory; (cf. Asplund et al. 2004). While this possibility may be worth considering, it can not yet be regarded as reliable in our opinion, since it causes serious discrepancies between theory and observation in the solar interior model (see, e.g., Young 2005 and the references therein). Besides, some questionable points still remain in their line-formation treatment (see also Appendix 1 in Takeda \\& Honda 2005).} in spite of not a few supportive evidences\\footnote{ Actually, the apparent subsolar tendency in the photospheric abundances of comparatively young stars has often been reported; e.g., C/N/O in early B main-sequence stars (Gies \\& Lambert 1992, Kilian 1992, see also Nissen 1993); C/N/O/Si/Mg/Al in early B stars (Kilian 1994); [Fe/H] in superficial normal late B and A stars (Sadakane 1990); [Fe/H] of B stars from UV spectra (Niemczura 2003); O in supergiants (Luck \\& Lambert 1985; Takeda \\& Takada-Hidai 1998).} since it contradicts the conventional scenario of galactic chemical evolution (where elemental abundances are generally believed to increase with time), we tend to regard this tendency as real, which means that the gas metallicity actually {\\it decreased} in an elapse of time between the formation of our Sun ($\\sim 5\\times 10^{9}$~yr ago) and the formation of $\\sim 2~M_{\\odot}$ stars ($\\la 10^{9}$~yr ago). Of course, in order to make this hypothesis more convincing, a reasonable explanation has to be done why such a reduction of the gas metallicity had occurred against the intuitive chemical evolution picture of increasing metallicity. While one such interpretation might be the dilution of the metallicity caused by an substantial infall of metal-poor primordial galactic gas speculated by Takeda et al. (2008), further observations and extensive abundance analyses on a much larger number of rapidly-rotating A dwarfs (as well as evolved G giants) would be required until we can say something about it with confidence." }, "0808/0808.3201_arXiv.txt": { "abstract": "We present sensitive, high angular resolution ($0\\rlap.{''}05$) VLA continuum observations made at 7 mm of the core of the HH~111/121 quadrupolar outflow. We estimate that at this wavelength the continuum emission is dominated by dust, although a significant free-free contribution ($\\sim$30\\%) is still present. The observed structure is formed by two overlapping, elongated sources approximately perpendicular to each other as viewed from Earth. We interpret this structure as either tracing two circumstellar disks that exist around each of the protostars of the close binary source at the core of this quadrupolar outflow or a disk and a jet perpendicular to it. Both interpretations have advantages and disadvantages, and future high angular resolution spectroscopic millimeter observations are required to favor one of them in a more conclusive way. ", "introduction": "It is generally accepted that most stars form in binary or multiple systems (Lada \\& Lada 2003). Furthermore, in the case of low and intermediate-mass stars it is also known that the process occurs with the presence of an accretion disk and a collimated outflow. From these two facts it follows that binary disk-jet systems should be common in regions of low-mass star formation. However, in practice these systems are hard to detect and identify and only a few have been studied in detail (e. g. Rodr\\'\\i guez et al. 1998; Anglada et al. 2004; Monin et al. 2007). Furthermore, it is still unclear if the members of a binary system will both be able to maintain the disks and outflows that characterize the formation of single stars. For example, the binary stars that form L1551~IRS5 are both believed to possess disks and jets (Rodr\\'\\i guez et al. 2003; Lim \\& Takakuwa 2006), while it is known that only one of the stars that forms the SVS 13 close binary system is associated with detectable circumstellar dust emission that is probably tracing a disk (Anglada et al. 2004). One of the most interesting cases of a binary source that is known to exhibit independent outflows, most probably associated with each of the components of the binary, is HH~111 (Reipurth et al. 1999). Located in Orion at a distance of 414 pc (Menten et al. 2007), the optical HH 111 jet was discovered by Reipurth (1989). It is an extremely well-collimated jet, aligned approximately in the east-west direction (at a PA of $\\sim 97^\\circ$) and whose knots move in the plane of the sky with velocities of the order of several hundred km s$^{-1}$ (Reipurth, Raga \\& Heathcote 1992). Reipurth, Bally \\& Devine (1997) found that this optical jet is part of a giant HH complex extending over 7.7 pc. This giant HH complex is very straight, suggesting great stability over the $10^4$ years of its lifetime. Near-infrared observations (Gredel \\& Reipurth 1993; 1994) revealed a second bipolar flow, named HH 121, that emerges from about the same position as the optical outflow, and is aligned approximately in the north-south direction (at a PA of $\\sim 35^\\circ$). This result suggested the presence of a close binary source in this region. Both the optical and the infrared outflows are detected as bipolar molecular outflows (Cernicharo \\& Reipurth 1996; Nagar et al. 1997; Lefloch et al. 2007). At the center of the quadrupolar outflow is the source IRAS 05491+0247 = VLA 1, a suspected class I binary with a total luminosity of about 25 L$_\\odot$ (e.g. Stapelfeldt \\& Scoville 1993, Yang et al. 1997). To advance in our understanding of this source a high angular resolution image at millimeter wavelengths was needed to compare with the information available for the quadrupolar outflow. ", "conclusions": "Our main conclusions follow: 1) Our high angular resolution ($0\\rlap.{''}05$) VLA continuum 7 mm observations of the core of the HH~111/121 quadrupolar outflow reveal a structure that can be described as two overlapping, elongated sources that appear approximately perpendicular to each other in the plane of the sky. 2) We discuss possible interpretations for this structure and conclude that the most viable ones are that we are observing two orthogonal disks around separate protostars or a disk with a perpendicular jet. Both intepretations have advantages and disadvantages, and high angular resolution spectroscopic millimeter observations (possible only in the future with the Atacama Large Millimeter Array) are required to disentangle what is going on at the core of this quadrupolar outflow. \\ack We thank an anonymous referee for valuable suggestions. LFR acknowledges the support of CONACyT, M\\'exico and DGAPA, UNAM. JMT and GA are supported by the MEC AYA2005-05823-C03 grant (co-funded with FEDER funds). GA also acknowledges support from Junta de Andaluc\\'{\\i}a." }, "0808/0808.0344_arXiv.txt": { "abstract": "The study of extragalactic sources of high energy radiation via the direct measurement of the proton and neutrino fluxes that they are likely to emit is one of the main goals for the future observations of the recently developed air showers detectors and neutrino telescopes. In this work we discuss the relation between the inclusive proton and neutrino signals from the ensemble of all sources in the universe, and the ``resolved'' signals from the closest and brightest objects. We also compare the sensitivities of proton and neutrino telescopes and comment on the relation between these two new astronomies. ", "introduction": "There is a general consensus that the highest energy cosmic rays (CR) are of extragalactic origin because they reach the Earth with an approximately isotropic angular distribution. The magnetic fields of the Milky Way are not sufficiently strong and extended to randomize the directions of particles produced by our own Galaxy, and the this isotropy of these CR is likely to reflect the large scale homogeneity of the universe. It is natural to expect that the extragalactic Ultra High Energy Cosmic Rays (UHECR) are produced in ``point--like'' astrophysical sources. The identification of these sources, and the clarification of the mechanisms that accelerate particles to these very large energies is clearly one of the crucial goals of the new large acceptance air shower detectors like the Pierre Auger Observatory. The trajectories of charged particles is bent by the presence of astrophysical magnetic fields, however at sufficiently large rigidity $E/Z$ (with $Z$ the particle electric charge) the magnetic deviations should become sufficiently small to allow the direct imaging of sources with cosmic rays. Unfortunately the intensity and structure of the intergalactic magnetic field are very poorly known, and therefore the region (in source distance and particle energy) where CR astronomy is possible remains very uncertain. It is possible (and there are observational hints) that the CR above $E \\gtrsim {\\rm few} \\times 10^{19}$~eV are already propagating in quasi--linear mode for source distances of order 100~Mpc or more, however the identification of the sources remains difficult because of the very small number of events available, with most of the sources contributing with not more than a single event. The analysis of the ``clustering'' of the observed events can give information about the luminosity of the individual sources. After the data of AGASA \\cite{Hayashida:1996bc,Takeda:1999sg} gave hints of possible clustering. this question has received a significant amount of attention \\cite{Uchihori:1999gu,Dubovsky:2000gv,Sommers:2008ji}. In this work we want to make a more quantitative and detailed study on how to interpret the results on the clustering. The nature of the particles in the UHECR remains a central open questions in the field. In this work we will assume that most of these particles are protons. This hypothesis is consistent with the existing data (if one takes into account the systematic uncertainties on the modeling of hadronic showers) and it is also favored in most theoretical models. Here this assumption is also made to allow a detailed quantitative description of particle propagation in intergalactic space, which is essential for the problem we are considering. It is possible and reasonably straightforward to generalize the discussion to the case of nuclei. All cosmic ray sources are unavoidably also sources of neutrinos \\cite{Gaisser:1994yf,Learned:2000sw} because some fraction of a population of relativistic hadrons will necessarily interact with ordinary matter or radiation fields targets inside or near their acceleration site creating pions and other weakly decaying particles that can (chain) decay into neutrinos. Viceversa, the emission of neutrinos imply the existence of relativistic hadrons, and therefore the acceleration of cosmic rays. The catalogues of the CR and neutrino sources are in principle identical. In practice the situation could more complicated, because the energy range studied by the CR and neutrino telescopes differ by several orders of magnitude, and the relation between the neutrino and cosmic ray emission from a source can vary significantly depending on the structure of the source, and it is certainly possible to have bright CR sources that emit a relatively small amount of neutrinos, and viceversa. Nonetheless, it is reasonable to expect that the brighest extragalactic sources of both CR and neutrinos will coincide, and it is interesting to discuss in parallel these new astronomies. Also in the case of neutrinos one can measure an inclusive flux, that sums the contributions from all sources in the universe, while the brightest and most powerful neutrino sources should be identifiable as individual objects. Therefore also for neutrinos the ``clustering'' of the detected events is a useful method of study. An important advantage is that for neutrinos linear propagation is certain, but one has to deal with the existence of the foreground of atmospheric neutrinos. The theoretical framework needed to discuss the clustering of neutrino events is essentially identical to the one needed for protons. A quantitative difference is that for neutrinos extragalactic space is perfectly transparent, and the inclusive flux receives most of its contribution from very distant and very faint sources, and the fraction of this flux that can be ``resolved'' in the contribution of identified sources is likely to be small. More that the interpretation of the existing data, the goal of this work is the development of some general analysis instruments that can be used in the study of future, higher statistics results. This work is organized as follows: in the next section we make a preliminary discussion of the so called ``Olbers Paradox''. The discussion of this celebrated puzzle allows to introduce the key concepts needed to compute the signal from the ensemble of all sources in the universe. In section~3 we collect the results on particle propagation in intergalactic space that are needed in the following. Section~4 discusses how the flux received from one astrophysical source depends on its redshift. Section~5 discusses possible forms of the luminosity function of the proton and neutrino sources. Section~6 compute the inclusive particle flux from the combined emission of all sources in the universe. Section~7 discusses what part of the inclusive flux can be resolved in the contribution of individual sources. Section~8 applies these analysis instruments to the interpretation of the recent Auger data. Section~9 discusses extragalactic neutrino astronomy. Section~10 gives some conclusions. ", "conclusions": "" }, "0808/0808.1081_arXiv.txt": { "abstract": "A common belief about big-bang cosmology is that the cosmological redshift cannot be properly viewed as a Doppler shift (that is, as evidence for a recession velocity), but must be viewed in terms of the stretching of space. We argue that, contrary to this view, the most natural interpretation of the redshift is as a Doppler shift, or rather as the accumulation of many infinitesimal Doppler shifts. The stretching-of-space interpretation obscures a central idea of relativity, namely that it is always valid to choose a coordinate system that is locally Minkowskian. We show that an observed frequency shift in any spacetime can be interpreted either as a kinematic (Doppler) shift or a gravitational shift by imagining a suitable family of observers along the photon's path. In the context of the expanding universe the kinematic interpretation corresponds to a family of comoving observers and hence is more natural. ", "introduction": "\\label{sec:intro} Many descriptions of big-bang cosmology declare that the observed redshift of distant galaxies is not a Doppler shift but is due to the ``stretching of space.'' The purpose of this paper is to examine the meaning of such statements and to assess their validity. We wish to make clear at the outset that we are not suggesting any doubt about either the observations or the general-relativistic equations that successfully explain them. Rather, our focus is on the interpretation: given that a photon does not arrive at the observer conveniently labeled ``Doppler shift,'' ``gravitational shift,'' or ``stretching of space,'' when can or should we apply these labels? Arguably an enlightened cosmologist never asks this question. In the curved spacetime of general relativity, there is no unique way to compare vectors at widely separated spacetime points, and hence the notion of the relative velocity of a distant galaxy is almost meaningless. Indeed, the inability to compare vectors at different points is the definition of a curved spacetime.\\cite{baezbunn,carroll,schutz,mtw} In practice, however, the enlightened view is far from universal. The view presented by many cosmologists and astrophysicists, particularly when talking to nonspecialists, is that distant galaxies are ``really'' at rest, and that the observed redshift is a consequence of some sort of ``stretching of space,'' which is distinct from the usual kinematic Doppler shift. In these descriptions, statements that are artifacts of a particular coordinate system are presented as if they were statements about the universe, resulting in misunderstandings about the nature of spacetime in relativity. In this paper we will show that the redshifts of distant objects in the expanding universe may be viewed as kinematic shifts due to relative velocities, and we will argue that if we are forced to interpret the redshift, this interpretation is more natural than any other. We begin with examples of the description of the cosmological redshift in the first three introductory astronomy textbooks chosen at random from the bookshelf of one of the authors. \\begin{itemize} \\item The cosmological redshift ``is {\\it not} the same as a Doppler shift. Doppler shifts are caused by an object's {\\it motion through space}, whereas a cosmological redshift is caused by the {\\it expansion of space}.''\\cite{kaufmann} (Emphasis in original.) \\item ``A more accurate view [than the Doppler effect] of the redshifts of galaxies is that the waves are stretched by the stretching of space they travel through \\ldots\\ If space is stretching during all the time the light is traveling, the light waves will be stretched as well.''\\cite{fraknoi} \\item ``Astronomers often express redshifts as if they were radial velocities, but the redshifts of the galaxies are not Doppler shifts \\ldots\\ Einstein's relativistic Doppler formula applies to motion through space, so it does not apply to the recession of the galaxies.''\\cite{seeds} \\end{itemize} More advanced textbooks often avoid this language. For instance, the books by Peacock\\cite{peacock} and Linder\\cite{linder} give particularly careful and clear descriptions of the nature of the cosmological redshift. However, statements similar to those we have cited can be found even in some advanced textbooks. For example, a leading advanced undergraduate level text states that Doppler shifts ``are produced by peculiar and not by recession velocities.''\\cite{harrison} In this paper we argue, as others have before us,\\cite{chodorowskimilne,chodorowski,peacocknotes,whiting} that statements such as these are misleading and foster misunderstandings about the nature of space and time. In general relativity the ``stretching of space'' explanation of the redshift is quite problematic. Light is governed by Maxwell's equations (or their general relativistic generalization), which contain no ``stretching of space term'' and no information on the current size of the universe. On the contrary, one of the most important ideas of general relativity is that spacetime is always locally indistinguishable from the (non-stretching) spacetime of special relativity, which means that a photon doesn't know about the changing scale factor of the universe.\\cite{footnote1} The emphasis in many textbooks on the stretching-of-spacetime interpretation of the cosmological redshift causes readers to take too seriously the stretching-rubber-sheet analogy for the expanding universe. For example, it is sometimes stated as if it were obvious that ``it follows that all wavelengths of the light ray are doubled'' if the scale factor doubles.\\cite{harrison} Although this statement is correct, it is not obvious. After all, solutions to the Schr\\\"odinger equation, such as the electron orbitals in the hydrogen atom, don't stretch as the universe expands, so why do solutions to Maxwell's equations? A student presented with the stretching-of-space description of the redshift cannot be faulted for concluding, incorrectly, that hydrogen atoms, the Solar System, and the Milky Way Galaxy must all constantly ``resist the temptation'' to expand along with the universe. One way to see that this belief is in error is to consider the problem sometimes known as the ``tethered galaxy problem,''\\cite{harrisontethered,davistethered} in which a galaxy is tethered to the Milky Way, forcing the distance between the two to remain constant. When the tether is cut, does the galaxy join up with the Hubble flow and start to recede due to the expansion of the universe? The intuition that says that objects suffer from a temptation to be swept up in the expansion of the universe will lead to an affirmative answer, but the truth is the reverse: unless there is a large cosmological constant and the galaxy's distance is comparable to the Hubble length, the galaxy falls toward us.\\cite{whiting,peacocknotes} Similarly, it is commonly believed that the Solar System has a very slight tendency to expand due to the Hubble expansion (although this tendency is generally thought to be negligible in practice). Again, explicit calculation shows this belief not to be correct.\\cite{sereno,cooperstock} The tendency to expand due to the stretching of space is nonexistent, not merely negligible. The expanding rubber sheet is quite similar to the ether of pre-relativity physics in that although it is intuitively appealing, it makes no correct testable predictions, and some incorrect ones such as the examples we have given. It therefore has no rightful place in the theory. (Some authors\\cite{barnes,francis} have argued that considerations such as these do not refute the notion that space is really expanding. We agree with the calculations in these papers but differ regarding the most useful language to use to describe the relevant phenomena.) In one set of circumstances the proper interpretation of the redshift seems clear. When the curvature of spacetime is small over the distance and time scales traveled by a photon, it is natural to interpret the observed frequency shift as a Doppler shift. This interpretation is the reason that a police officer can give you a speeding ticket based on the reading on a radar gun. As far as we know, no one has successfully argued in traffic court that there is an ambiguity in interpreting the observed frequency shift as a Doppler shift.\\cite{footnote3} In the expanding universe, spacetime curvature is small over regions encompassing nearby objects, specifically those with $z=\\Delta\\lambda/\\lambda\\ll 1$. There should be no hesitation about calling the observed redshifts Doppler shifts in this case, just as there is none in traffic court. Surprisingly, however, many people seem to believe that the ``stretching of space'' interpretation of the redshift is the only valid one, even in this limit. We will examine the interpretation of redshifts of nearby objects more carefully in Sec.~\\ref{sec:lowz}. Aside from low-redshift sources, there is another case in which spacetime curvature can be neglected in considering cosmological redshifts, namely low-density cosmological models. An expanding universe with density $\\Omega=0$ (often known as the Milne model\\cite{milne}) is merely the flat Minkowski spacetime of special relativity expressed in nonstandard coordinates.\\cite{footnote4} In an $\\Omega=0$ universe there are no gravitational effects at all, so any observed redshift, even of a very distant galaxy, must be a Doppler shift. Furthermore, for low but nonzero density ($\\Omega \\ll 1$), the length scale associated with spacetime curvature is much longer than the horizon distance. Hence, spacetime curvature effects (that is, gravitational effects) are weak throughout the observable volume, and the special-relativistic Doppler shift interpretation remains valid even for galaxies with arbitrarily high redshifts. The more interesting cases are when $z$ and $\\Omega$ are not small; that is, when the source is sufficiently distant that gravitational effects are important over the photon's trajectory. The consensus is that the Doppler shift language must be eschewed in this setting. In Sec.~\\ref{sec:highz} we review a standard argument\\cite{peacock,peacocknotes} that even in this case the redshift can be interpreted as the accumulation of infinitesimal Doppler shifts along the line of sight, and we further argue that there is a natural way to interpret the redshift as a single (non-infinitesimal) Doppler shift. A common objection to this claim is that the coordinate velocity is not related to the redshift in accordance with the special-relativistic Doppler formula.\\cite{davissuperluminal,davisconfusion} However, the velocity referred to in this claim is a mere artifact of a particular choice of time coordinate. Specifically, it is the rate of change of the distance to the object with respect to the cosmic time coordinate, as measured at the present cosmic time. This coordinate velocity is an unnatural quantity to discuss, because it depends on data outside of the observer's light cone. The more natural velocity is the velocity of the object at the time it crossed our past light cone, relative to us today. This velocity is also a coordinate-dependent concept, but as we will show in Sec.~\\ref{sec:highz}, the most natural way to specify it operationally\\cite{synge,narlikar} leads to a result that is consistent with the special-relativistic Doppler formula. In Sec.~\\ref{sec:observers} we widen our focus to consider frequency shifts in arbitrary curved spacetimes. In any curved spacetime the observed frequency shift in a photon can be interpreted as either a kinematic effect (a Doppler shift) or as a gravitational shift. The two interpretations arise from different choices of coordinates, or equivalently from imagining different families of observers along the photon's path. We will describe this construction explicitly, and show that the comoving observers who are usually used to describe phenomena in the expanding universe are the ones that correspond to the Doppler shift interpretation. ", "conclusions": "" }, "0808/0808.2611_arXiv.txt": { "abstract": "We present new evolution sequences for very low mass stars, brown dwarfs and giant planets and use them to explore a variety of influences on the evolution of these objects. While the predicted adiabatic evolution of luminosity with time is very similar to results of previous work, the remaining disagreements reveal the magnitude of current uncertainty in brown dwarf evolution theory. We discuss the sources of those differences and argue for the importance of the surface boundary condition provided by atmosphere models including clouds. The L- to T-type ultracool dwarf transition can be accommodated within the \\citet{am01} cloud model by varying the cloud sedimentation parameter. We develop a simple model for the evolution across the L/T transition. By combining the evolution calculation and our atmosphere models, we generate colors and magnitudes of synthetic populations of ultracool dwarfs in the field and in galactic clusters. We focus on near infrared color-magnitude diagrams (CMDs) and on the nature of the ``second parameter'' that is responsible for the scatter of colors along the $\\teff$ sequence. Instead of a single second parameter we find that variations in metallicity and cloud parameters, unresolved binaries and possibly a relatively young population all play a role in defining the spread of brown dwarfs along the cooling sequence. We also find that the transition from cloudy L dwarfs to cloudless T dwarfs slows down the evolution and causes a pile up of substellar objects in the transition region, in contradiction with previous studies. The same model is applied to the Pleiades brown dwarf sequence with less success, however. Taken at face value, the present Pleiades data suggest that the L/T transition occurs at lower $\\teff$ for lower gravity objects, such as those found in young galactic clusters. The simulated populations of brown dwarfs also reveal that the phase of deuterium burning produces a distinctive feature in CMDs that should be detectable in $\\sim 50$--100$\\,$Myr old clusters. ", "introduction": "There are now approximately 450 L dwarfs and 100 T dwarfs known (see \\citet{kirk05} for a review of these spectral classes). They span effective temperatures from about 2400 to 700 K and exhibit a range of gravities, metallicities, and atmospheric condensate contents. After more than a decade of intense study, the modeling of the complex atmospheres and synthetic spectra of brown dwarfs has reached a rather high degree of sophistication, including the chemistry of a very large number of gas and condensate species \\citep{allard01,lod02}, increasingly complete molecular opacity databases \\citep{fml08,sb07}, extreme resonance line broadening \\citep{bms00,aak05}, and particulate cloud models \\citep{allard01,am01,tsu02,helling08}. While a few conspicuous problems remain, the synthetic spectra and colors reproduce the observations fairly well and the determination of the basic astrophysical properties of brown dwarfs has begun in earnest. The full astrophysical benefit of synthetic spectra and colors is obtained when atmosphere calculations are coupled with evolution models that provide the surface atmospheric parameters $(\\teff,g)$ as a function of mass and age. The time evolution of the spectrum and colors, as well as absolute fluxes can be computed and directly compared with observations to estimate astrophysical parameters that are not easily amenable to direct observation. To enable such comparisons using our own model atmosphere effort and to pursue a more complete analysis of spectroscopic and photometric data, we have developed a code to compute evolution sequences of low mass stars, brown dwarfs and giant planets. These evolution sequences have already been applied extensively to the analysis of brown dwarf observations \\citep{roellig04,saumon06,saumon07,leggett07y, leggett07p,cushing08} but have not been discussed in any detail. Here, we describe the input physics and assumptions of our particular approach as well as the unique aspects of our atmospheric boundary condition. The evolution of isolated brown dwarfs---a relatively simple case of stellar evolution---has been studied extensively for the past 20 years and is well understood \\citep{dm85,nrj86,bhl89,bl93,bur97,cbah00a,cbah00b,bcah02}. Our evolution model is similar to the more recent work and the result are very similar to previous work. A detailed comparison with other published calculations of the evolution of brown dwarfs reveals small differences that we quantify and, to the extent possible, attribute to the different assumptions and approximations in each model. We highlight the application of the evolution sequences to the calculation of synthetic near infrared color-magnitude diagrams (CMDs) to explore the nature of the ``second parameter'' responsible for the spread of observed objects around the main trends along the L and T spectral sequences. We develop a simple parametric model for the L/T transition to reproduce with good success the CMD of field brown dwarfs. We apply the same model to synthesize the population of the Pleiades cluster (110$\\,$Myr) and compare with the latest deep survey data for this galactic cluster. We discuss several potentially observable features in near-infrared CMDs that would illuminate the evolution of brown dwarfs as well as the nature of L/T transition. ", "conclusions": "Our calculation of the evolution of very low mass stars, brown dwarfs and planetary mass objects produces models that are quantitatively in very good agreement with published calculations. A detailed comparison shows some systematic differences that can be attributed to conductive energy transport, different choices of composition for the interior, and for the initial state but the primary source of discrepancy is the surface boundary condition. The cloudless and cloudy model atmospheres from which we extract the surface boundary condition have been validated with extensive comparisons with spectroscopic and photometric data. Thus, our boundary condition is quite realistic and we do not expect that the foreseeable improvements in atmosphere models will have much effect on our modeled evolution of brown dwarfs. By using the boundary condition from our atmosphere models, we can compute self-consistently the evolution, absolute fluxes, absolute magnitudes, and colors for a variety of cloud properties, metallicities and eventually including vertical mixing \\citep{saumon06}. Only the evolution across the L/T transition region cannot be modeled in this self-consistent fashion because a transition cloud model is not yet available. We have developed a simple model for the cooling and color evolution of brown dwarfs across the L/T transition. This hybrid model predicts an excess of brown dwarfs in the $\\teff$ range of the transition by about a factor of 2 compared to purely cloudy or cloudless evolution. We have applied this hybrid evolution model, combined with the near-infrared magnitudes predicted by our atmosphere models, to generate synthetic CMDs that can be compared with samples of brown dwarfs in the field and in galactic clusters. Our primary focus is the ``second parameter'' responsible for the dispersion about the brown dwarf sequence in the CMD. Population synthesis is a potentially powerful tool but the results can be quite sensitive to the input assumptions for the IMF, SFR, metallicity, and the binary frequency and mass ratio distribution. Our knowledge of the brown dwarf population in the solar neighborhood cannot yet provide these inputs {\\it a priori}. The relatively small number of brown dwarfs with known parallax and the heterogeneous nature of the sample imply that the observed distribution of field substellar dwarfs in the CMD is not yet fully characterized. Nevertheless, both observations and models have reached a stage where general trends in the ``second parameter'' along the L-T spectral sequence can be interpreted. We find that for our fiducial assumptions (power law IMF with $\\alpha=1$, constant SFR over the past 10$\\,$Gyr, single brown dwarfs only and [M/H]=0), the hybrid sequence reproduces the overall sequence from late M through late T dwarfs rather well, but not the dispersion along the sequence. Based on the near-infrared CMDs, we find that the L/T transition occurs between $\\teff \\sim 1400-1200\\,$K in field brown dwarfs, in agreement with previous estimates. While a transition over such a narrow range of $\\teff$ appears to be ``fast'' considering the rather dramatic change in $JHK$ colors across the transition, these values of $\\teff$ correspond to ages of 2 and 4$\\,$Gyr for a 0.06$\\,M_\\odot$ brown dwarf. The duration of the transition decreases rapidly with mass however, lasting only 0.15$\\,$Gyr for a 0.03$\\,M_\\odot$ brown dwarf. Better agreement can be obtained from late M to late L spectral types if the population is younger, such as with a constant SFR that started only 5$\\,$Gyr ago, by including binaries, or assuming that there is a wider range of cloud properties for later L spectral types. For a fixed metallicity, all simulations predict that the distribution of brown dwarfs in the CMD will have a sharp edge formed by old brown dwarfs of all masses ($\\gtrsim 3\\,$Gyr). This edge is on the blue side of the distribution for $M_K \\lesssim 12.5$ and to the red side after the $J-K$ color of the sequence turns over, corresponding to $M_K \\gtrsim 13$. This feature is not visible in the data, however, most likely because it is blurred by variations in metallicity within the sample. We are not able to include metallicity variations in simulations of cloudy brown dwarfs, except in a very approximate way (Fig. \\ref{fig:metal_f2}). We find that it could be a significant contributor to the second parameter. Detailed spectral analysis of brown dwarfs with unusual $J-K$ colors for their spectral types have more extreme cloud parameters \\citep{burgasser08,cushing08,stephens08}, which strongly suggest that cloudiness is the second parameter. Spectral analysis with models of non-solar metallicities have barely begun, however, so it would be premature to attribute the dispersion along the L sequence entirely to cloud characteristics. On the other hand, a simulation of cloudless models with an empirical metallicity distribution shows a good match to the dispersion of the late T dwarfs. For those coolest dwarfs, we find that the second parameter is a combination of metallicity variations (which dominate) and binaries. Gravity is not important as the with of the distribution is not affected by the choice of SFR or IMF. To summarize, we find that there is no single second parameter that accounts for the dispersion of brown dwarfs around the sequence seen in the CMD. A young age distribution, a range of metallicities and cloud properties as well as binaries all contribute to the dispersion. The challenge will be to untangle their contributions. The hybrid model fares somewhat worse when compared to the much younger brown dwarf population of the Pleiades. If the two faintest Pleiads reported are indeed T dwarf members of the cluster, then they provide strong evidence that the L/T transition occurs at lower $\\teff$ in lower gravity objects (i.e. younger or less massive). Finally, isochrones in CMDs clearly reveal the phase of deuterium burning at young ages, a feature that should be observable in young clusters with ages between of $\\sim 50-100\\,$Myr. At this time, the modest size of the sample of L and T dwarfs with known parallax and the lingering problems in modeling the atmospheres of cloudless and cloudy brown dwarfs restrict how much we can learn from the study of CMDs. Model limitations will eventually be overcome as new moelcular line lists are being developed for key molecules and cloud models become more sophisticated. We anticipate a rich harvest of brown dwarf parallaxes from the volume limited solar neighborhood census component of the Panoramic Survey Telescope \\& Rapid Response System (Pan-STARRS) \\footnote{\\tt http://pan-starrs.ifa.hawaii.edu/project/reviews/PreCoDR/documents/scienceproposals/sol.pdf}. Color-magnitude diagrams are a potentially powerful tool for the study of brown dwarf evolution and of the L/T transition. Statistical comparisons with synthetic populations in two-dimensional parameter space will become an important complement to the detailed studies of the spectra of individual transition objects and of brown dwarf binaries." }, "0808/0808.0614_arXiv.txt": { "abstract": "Fermionic condensate and the vacuum expectation values of the energy-momentum tensor are investigated for twisted and untwisted massive spinor fields in higher-dimensional de Sitter spacetime with toroidally compactified spatial dimensions. The expectation values are presented in the form of the sum of corresponding quantities in the uncompactified de Sitter spacetime and the parts induced by non-trivial topology. The latter are finite and renormalizations are needed for the first parts only. Closed formulae are derived for the renormalized fermionic vacuum densities in uncompactified odd-dimensional de Sitter spacetimes. It is shown that, unlike to the case of 4-dimensional spacetime, for large values of the mass, these densities are exponentially suppressed. Asymptotic behavior of the topological parts in the expectation values are investigated in the early and late stages of the cosmological expansion. When the comoving lengths of compactified dimensions are much smaller than the de Sitter curvature radius, the leading term in the topological parts coincide with the corresponding quantities for a massless fermionic field and are conformally related to the corresponding flat spacetime results. In this limit the topological parts dominate the uncompactified de Sitter part and the back-reaction effects should be taken into account. In the opposite limit, for a massive field the asymptotic behavior of the topological parts is damping oscillatory. ", "introduction": "De Sitter (dS) spacetime is one of the simplest and most interesting spacetimes allowed by general relativity. Quantum field theory in this background has been extensively studied during the past two decades. Much of early interest to dS spacetime was motivated by the questions related to the quantization of fields propagating on curved backgrounds. This spacetime has a high degree of symmetry and numerous physical problems are exactly solvable on this background. The importance of this theoretical work increased by the appearance of \\ the inflationary cosmology scenario \\cite% {Lind90}. In most inflationary models, an approximately dS spacetime is employed to solve a number of problems in standard cosmology. During an inflationary epoch, quantum fluctuations in the inflaton field introduce inhomogeneities and may affect the transition toward the true vacuum. These fluctuations play a central role in the generation of cosmic structures from inflation. More recently astronomical observations of high redshift supernovae, galaxy clusters and cosmic microwave background \\cite{Ries07} indicate that at the present epoch, the Universe is accelerating and can be well approximated by a world with a positive cosmological constant. If the Universe would accelerate indefinitely, the standard cosmology would lead to an asymptotic dS universe. Hence, the investigation of physical effects in dS spacetime is important for understanding both the early Universe and its future. Many of high energy theories of fundamental physics are formulated in higher-dimensional spacetimes. In particular, the idea of extra dimensions has been extensively used in supergravity and superstring theories. It is commonly assumed that the extra dimensions are compactified. From an inflationary point of view, universes with compact spatial dimensions, under certain conditions, should be considered a rule rather than an exception \\cite{Lind04}. The models of a compact universe with non-trivial topology may play important roles by providing proper initial conditions for inflation. As it was argued in Refs. \\cite{McIn04}, there is no reason to believe that the version of dS spacetime which may emerge from string theory, will necessarily be the most familiar version with symmetry group $% O(1,4)$ and there are many different topological spaces which can accept the dS metric locally. There are many reasons to expect that in string theory the most natural topology for the universe is that of a flat compact three-manifold. The quantum creation of the universe having toroidal spatial topology is discussed in \\cite{Zeld84} and in references \\cite{Gonc85} within the framework of various supergravity theories. The compactification of spatial dimensions leads to a number of interesting quantum field theoretical effects which include instabilities in interacting field theories \\cite{Ford80a}, topological mass generation \\cite{Ford79} and symmetry breaking \\cite{Toms80b}. In the case of non-trivial topology, the boundary conditions imposed on fields give rise to the modification of the spectrum for vacuum fluctuations and, as a result, to the Casimir-type contributions in the vacuum expectation values of physical observables (for the topological Casimir effect and its role in cosmology see \\cite% {Grib94,Most97} and references therein). In the Kaluza-Klein-type models, the Casimir effect has been used as a stabilization mechanism for moduli fields which parametrize the size and the shape of the extra dimensions. The Casimir energy can also serve as a model for dark energy needed for the explanation of the present accelerated expansion of the universe (see \\cite% {Milt03} and references therein). One-loop quantum effects for various spin fields on the background of dS spacetime, have been discussed by several authors (see, for instance, \\cite{Cher68}-\\cite{Birr82} and references therein). The effects of the toroidal compactification of spatial dimensions in dS spacetime on the properties of quantum vacuum for a scalar field with general curvature coupling parameter are investigated in Refs. \\cite% {Saha07,Bell08} (for quantum effects in braneworld models with dS spaces and in higher-dimensional brane models with compact internal spaces see, for instance, Refs. \\cite{dSbrane,Flac03}). The one-loop quantum effects for a fermionic field on background of 4-dimensional dS spacetime with spatial topology $\\mathrm{R}^{p}\\times (\\mathrm{S}^{1})^{q}$ are studied in \\cite% {Saha08}. In the present paper, we investigate one-loop quantum effects arising from vacuum fluctuations of a fermionic field on background of higher-dimensional dS spacetime with toroidally compactified spatial dimensions. The important quantities that characterize the quantum fluctuations during the dS expansion are the fermionic condensate and the expectation value of the energy-momentum tensor. In the next section, by using the dimensional regularization procedure, we evaluate these quantities in uncompactified odd-dimensional dS spacetimes. The plane wave fermionic eigenfunctions in $% (D+1)$-dimensional dS spacetime with an arbitrary number of toroidally compactified dimensions are constructed in section \\ref{sec:EigFunc}. In section \\ref{sec:FermCond} these eigenfunctions are used for the evaluation of the fermionic condensate in both cases of the fields with periodicity and antiperiodicity conditions along compactified dimensions. The behavior of these quantities are investigated in asymptotic regions of the parameters. The topological parts in the vacuum expectation values of the energy-momentum tensor are investigated in section \\ref{sec:EMT}. In the last section we summarize the main results of the paper. ", "conclusions": "\\label{sec:Conc} In the present paper we have investigated the fermionic condensate and the VEV of the energy-momentum tensor for a massive fermionic field in higher-dimensional dS spacetime with toroidally compactified spatial dimensions. In Section \\ref{sec:UncompdS} we have considered the corresponding quantities in uncompactified odd-dimensional dS spacetime assuming that the field is prepared in the Bunch-Davies vacuum state. The renormalization is done by using the dimensional regularization procedure. Closed expressions, formulae (\\ref{FermConddSRen}) and (\\ref{TkldSren}), are derived for the renormalized fermionic condensate and the VEV of the energy-momentum tensor respectively. For large values of the mass these quantities are exponentially suppressed. Note that in even-dimensional dS spacetime for large mass the suppression is power-law. Further, we have investigated one-loop quantum effects on the fermionic vacuum induced by the non-trivial topology of spatial dimensions. Specifically, we have considered the dS spacetime with toroidally compactified dimensions having the spatial topology $\\mathrm{R}^{p}\\times (% \\mathrm{S}^{1})^{q}$. For the evaluation of the vacuum densities, the mode-summation procedure is employed. In this procedure we need to know the corresponding eigenspinors satisfying appropriate boundary conditions along the compactified dimensions. These eigenspinors are constructed in section % \\ref{sec:EigFunc} for both fields obeying periodicity and antiperiodicity boundary conditions. By using these eigenfunctions and applying to the mode-sums the Abel-Plana formula, the VEVs for the spatial topology $\\mathrm{% R}^{p}\\times (\\mathrm{S}^{1})^{q}$ are presented in the form of the sum of the corresponding quantity in the topology $\\mathrm{R}^{p+1}\\times (\\mathrm{S% }^{1})^{q-1}$ and of the part which is induced by the compactness of $(p+1)$% th dimension. For fields obeying periodicity conditions, the topological parts are given by formulae (\\ref{DeltCond2}) and (\\ref{TopTll}) for fermionic condensate and energy-momentum tensor, respectively. The corresponding formulae for the field with antiperiodicity conditions are obtained from those for the field obeying periodicity conditions inserting the factor $(-1)^{n}$ in the summation over $n$ and replacing the definition for $k_{\\mathbf{n}_{q-1}}^{2}$ by (\\ref{knq-1Tw}). The topological parts are finite and the renormalization procedure is needed only for the uncompactified dS spacetime. These parts are time-dependent and break the dS symmetry. The corresponding vacuum stresses along the uncompactified dimensions coincide with the energy density and, hence, in the uncompactified subspace the equation of state for the topological part of the energy-momentum tensor is of the cosmological constant type. For a massless fermionic field the problem under consideration is conformally related to the corresponding problem in the Minkowski spacetime with spatial topology $\\mathrm{R}^{p}\\times (\\mathrm{S}^{1})^{q}$ and the topological part of the fermionic condensate vanishes. For the VEV of the energy-momentum tensor we have the standard relation $\\langle T_{k}^{l}\\rangle _{c}=a^{-(D+1)}(\\eta )\\langle T_{k}^{l}\\rangle _{c}^{% \\mathrm{(M)}}$ between the topological contributions. For a massive fermionic field, in the limit when the comoving length of a compactified dimension is much smaller than the dS curvature radius, the topological part in the VEV of the energy-momentum tensor coincides with the corresponding quantity for a massless field and is conformally related to the VEV in toroidally compactified Minkowski spacetime. In particular, the topological part in the vacuum energy density is positive for an untwisted fermionic field. This limit corresponds to the early stages of the cosmological evolution and the topological parts dominate over the uncompactified dS parts. At these stages the back-reaction effects of the topological terms are important and these effects can essentially change the dynamics of the model. In the opposite limit, when the comoving lengths of the compactified dimensions are large with respect to the dS curvature radius, in the case of a massive field the asymptotic behavior of the topological parts are oscillatory damping for both fermionic condensate and the energy-momentum tensor and their respectively leading term are given by formulae (\\ref{CondLate}) and (\\ref{TllSmall}). These formulae describe the behavior of the topological parts in the late stages of the cosmological expansion. As the corresponding uncompactified dS parts are time-independent, we have similar oscillations in the total VEVs as well. Note that this type of oscillatory behavior is absent for a massless fermionic field." }, "0808/0808.0939_arXiv.txt": { "abstract": "A general phenomenological theory is presented for the phase behavior of ferromagnetic superconductors with spin-triplet electron Cooper pairing. The theory accounts in detail for the temperature-pressure phase diagram of ZrZn$_2$, while the main features of the diagram for UGe$_2$ are also described. Quantitative criteria are deduced for the U-type (type I) and Zr-type (type II) unconventional ferromagnetic superconductors with spin-triplet Cooper electron pairing. Some basic properties of quantum phase transitions are also elucidated. ", "introduction": " ", "conclusions": "" }, "0808/0808.0422_arXiv.txt": { "abstract": "{Below 1 mHz, the power spectrum of helioseismic velocity measurements is dominated by the spectrum of convective motions (granulation and supergranulation) making it difficult to detect the low-order acoustic modes and the gravity modes.} {We want to better understand the behavior of solar granulation as a function of the observing height in the solar atmosphere and with magnetic activity during solar cycle 23. } {We analyze the Power Spectral Density (PSD) of eleven years of GOLF/SOHO velocity-time series using a Harvey-type model to characterize the properties of the convective motions in the solar oscillation power spectrum. We study then the evolution of the granulation with the altitude in the solar atmosphere and with the solar activity.} {First, we show that the traditional use of a lorentzian profile to fit the envelope of the $p$ modes is not well suitable for GOLF data. Indeed, to properly model the solar spectrum, we need a second lorentzian profile. Second, we show that the granulation clearly evolves with the height in the photosphere but does not present any significant variation with the activity cycle.} {} ", "introduction": "\\label{Intro} The photosphere, the visible layer of the Sun, is the location where the energy transport previously dominated by convection and turbulence is largely done by radiation with an optical depth of $2/3$. The gas is visible in the form of granules, that penetrate inside the stable photosphere. These granules, as well as other larger structures like the mesogranules or the supergranules, are the manifestation of the different spatial scales of the convective motions occurring in this region of the Sun \\citep{Zahn87,Roudier91,Espagnet93}. The study of the granulation is particularly important in helioseismology because, on the one hand, it excites the so-called 5-min oscillations, i.e. the acoustic (p) modes, and, on the other hand, it dominates the power spectrum at low frequencies preventing the detection of low-order p modes. Indeed, in the case of velocity measurements, the lower detection limit of acoustic modes is established around 1 mHz \\citep{Garcia01,Garcia04a,Broomhall07}. To progress in the detection of such modes as well as to increase the detection probability of gravity modes \\citep{Appourchaux00,Gabriel02,Turck04,Garcia07,Mathur07} we need to better characterize the properties of the granulation in order to reduce, if possible, their impact on the helioseismic measurements. The coming GOLF-NG instrument will soon address this problem \\citep{Turck06}. This new-generation instrument should improve the signal-to-noise (S/N) ratio of low-frequency modes by measuring the Doppler velocity at different heights in the solar atmosphere. It would thus benefit from the reduction in the coherence of the granulation with the atmospheric altitude \\citep{Garcia04b}. In this paper, we analyze the Power Spectral Density (PSD) of velocity time sub-series from the GOLF\\footnote{Global Oscillation at Low Frequencies \\citep{Gabriel95}} instrument on board SOHO\\footnote{SOlar and Heliospheric Observatory \\citep{Domingo95}} which is a solar disk integrated resonant spectrometer. With such an instrument, we can already study the mean behaviour of the solar granulation in some range of the atmosphere. So this work comes in complement to the study of \\citet{Espagnet95} where the authors found that the photosphere is highly structured with two distinct layers below and above about 90 km. With GOLF, and the technique used in \\citet{Jimenez07}, we are able to study, without spatial resolution, a region located between 250 up to 550 km above the photosphere. We show that the granulation evolves with the height in the photosphere and that the granules tend to have shorter lifetimes with a weaker velocity when higher in the atmosphere. Section \\ref{Analysis} is devoted to the data analysis, with first a brief summary of the velocity calibration procedure of the GOLF signal and secondly a description of the fitting procedure of the power spectra in more details. Section \\ref{Results} is dedicated to a detailed study of the granulation motions and its evolution with time during the solar cycle. Finally, last section (Sect. \\ref{Discussion}) will emphasize the main results of this paper and will anticipate on future incoming works. ", "conclusions": "\\label{Discussion} In this paper we have investigated the vertical structure and time evolution of the solar granulation by means of a novel methodology based on the analysis of the full-disk Sun-as-a-star Doppler velocity observations. Thus we have been able to study the vertical velocity fluctuations and lifetimes of the solar granulation. We have shown that the GOLF PSD can be correctly characterized by our model, a Harvey function with two Lorentzian profiles. This work extends the study of \\citet{Espagnet95} where they found that the photosphere is highly structured with two distinct layers below and above about 90 km. With GOLF, we study the photosphere above $\\approx$ 280 km and we showed that granules tend to live longer with a weaker velocity when higher in the atmosphere. Following the results of \\citet{Title89}, there is a strong correlation between the granule sizes and lifetime. Therefore, we can conclude that larger granules reach the top of the photosphere, while penetration of small granules decreases with the size. To be more precise, as Figure \\ref{fig3} shows a lifetime of about 400-550 s for granules between 250 and 550 km, we can estimate from the article of \\citet{Title89} and their figure 21 that the granules in these altitudes have a lifetime-average size of about 1.2 arcsec and a maximum size of about 1.4-1.5 arcsec. We have found that the granulation rms velocities ($\\sigma_{gr}$) are between 0.28 and 0.4 $ms^{-1}$. These values are very different from those obtained from high-resolution measurements which are about 1 $kms^{-1}$. The difference in spatial resolution is the most likely cause of this difference of three orders of magnitude. However we did not put in evidence a change with the solar activity. This is consistent with the result of \\citet{Jimenez03} that the solar cycle effects are very small compared to the change in the observing height in the photosphere due to the orbital motion. It could be due to the fact that GOLF observes in the more stable part of the sodium lines. However if we do not find any significant variation with the cycle, an other study \\citep{Muller07} found a cyclic variation of the contrast of the granules, nearly in phase with the solar cycle, the contrast being smaller at the periods of solar maximum, but no corresponding variation in the scale. This work opens two perspectives: (1) to better understand the evolution of the whole solar convective background during the activity cycle, and specially the evolution of the acoustic-mode envelope presently characterized by the presence of two Lorentzians. We are likely to think that the excess power characterized by the second Lorentzian corresponds to the presence of chromospheric modes, (2) to extend this study to our new instrument GOLF-NG observation. It will cover a larger part of the solar atmosphere thanks to the 8 extractions of the Doppler velocity with a proper determination of their location due to the measurement of both wings and the use of only the D1 line \\citep{Jimenez07, Turck06, Turck08}." }, "0808/0808.0570_arXiv.txt": { "abstract": " ", "introduction": "\\label{sec:1} The original classification of elliptical galaxies, which goes back to Edwin Hubble, is based on their apparent ellipticity. The projected ellipticity, however, is also a function of inclination, and not only of the intrinsic ellipticity. \\cite{B88} showed that the isophotes of elliptical galaxies deviate significantly from perfect ellipses. The deviations can be quantified by the $a_4$-parameter, where negative $a_4$ values signify boxy deviations and positive $a_4$ values discy deviations from a perfect ellipse. But it was also found that ellipticals which have boxy isophotes, are also X-ray bright, more luminous, rotate slowly and have cored inner surface density slopes, while discy ellipticals are less luminous, rotate fast and have power-law surface density profiles \\cite{BDM88}. This led \\cite{KB96} to revise the Hubble classification of ellipticals using isophotal shape instead of ellipticity as a galaxy family membership criterium. One of the most successful models of elliptical galaxy formation is the merging of disc galaxies. \\cite{BA98} proposed that the isophotal shape depends on the violence of the merging and subsequently \\cite{N99} showed that mergers of galaxies with comparable mass form more likely boxy ellipticals, while mergers of galaxies with differing mass form discy ellipticals. However, disc-disc mergers do not form a perfect dichtomy \\cite{N03}, while elliptical-elliptical mergers seem to disconnect merging ratio and isophotal shape, i.e. they are always boxy \\cite{NKB}. In contrast to this gas seems to play an important role in fast-rotating, discy ellipticals \\cite{N06},\\cite{J07}. However, recently the SAURON sample, a survey of 48 elliptical and lenticular galaxies, \\cite{E04} showed that isophotal shapes are not well connected to rotation, i.e. fast rotating galaxies can have boxy isophotes \\cite{E07}. In the following we want to shed some light on the connection of photometric and kinematic parameters in N-body merger remnants and how they relate to the internal orbital structure. ", "conclusions": "The trends seen here are typical for disc-disc merger remnants, while they are probably representative for low to intermediate luminosity elliptical galaxies, they cannot explain the whole population of elliptical galaxies. The analysis of the orbital fine structure should be extended to a broader range of formation mechanisms, e.g. E/S0 galaxies with boxy isophotes, will certainly have bars \\cite{A05} which have not been covered by our present analysis. Dry merging will have certainly played a role in the formation of the most massive stellar systems in the universe and will lead to round or boxy systems \\cite{NKB},\\cite{KB05}, while gas physics is important to form remnants with realistic LOSVDs \\cite{N06},\\cite{J07}. Monolithic collapse can also form triaxial self-gravitating systems which can have higher fractions of semi-stochastic orbits than our remnants \\cite{Aqui07}. Finally binary merger remnants can be compared to elliptical galaxies which formed in cosmological simulations and will probably have very different orbital structures\\cite{N07}." }, "0808/0808.0093_arXiv.txt": { "abstract": "{Despite many studies of star formation in spiral galaxies, a complete and coherent understanding of the physical processes that regulate the birth of stars has not yet been achieved, nor has unanimous consensus been reached, despite the many attempts, on the effects of the environment on the star formation in galaxy members of rich clusters.} {We focus on the local and global Schmidt law and we investigate how cluster galaxies have their star formation activity perturbed.} {We collect multifrequency imaging for a sample of spiral galaxies, members of the Virgo cluster and of the local field; we compute the surface density profiles for the young and for the bulk of the stellar components, for the molecular and for the atomic gas.} {Our analysis shows that the bulk of the star formation correlates with the molecular gas, but the atomic gas is important or even crucial in supporting the star formation activity in the outer part of the disks. Moreover, we show that cluster members that suffer from a moderate HI removal have their molecular component and their SFR quenched, while highly perturbed galaxies show an additional truncation in their star forming disks.} {Our results are consistent with a model in which the atomic hydrogen is the fundamental fuel for the star formation, either directly or indirectly through the molecular phase; therefore galaxies whose HI reservoirs have been depleted suffer from starvation or even from truncation of their star formation activity.} ", "introduction": "A satisfactory understanding of the physical processes governing the formation of giant molecular clouds (GMC) from primeval atomic hydrogen (HI) and their instability leading to the formation of stars is still far from achieved. This is not surprising due to the complex mix of hydrodynamical and gravitational physics involved in the process of star formation. Beside the theoretical difficulties, what surprisingly limits our present understanding of this issue is the lack of a robust phenomenology of star formation in nearby galaxies. Despite the many recipes that have been proposed\\footnote{See for instance the dynamically modified Schmidt law in \\citet{won02}}, the simplest and most commonly used parametrization is the Schmidt law \\citep{sch59} that establishes a correlation between the star formation rate density $\\rho_{sfr}$ and the gas density $\\rho_{gas}$: \\begin{equation} \\rho_{sfr}=a\\rho_{gas}^\\alpha\\:; \\end{equation} derived for our Galaxy, the validity of this law has been tested also for external galaxies \\citep[e.g.][]{ken83}, for which a similar relation, usually written in terms of surface densities, holds: \\begin{equation}\\label{Slaw} \\Sigma_{sfr}=A\\Sigma_{gas}^n\\:. \\end{equation} Many studies of the global Schmidt law, i.e. the relationship between the disk--averaged star formation rate (SFR) and the total gas content, have been carried out on large samples of galaxies. A milestone paper on this issue is by \\citet{ken98}, who studied the connection between the integrated SFR surface density and the integrated gas surface density in a sample of normal and starburst galaxies. Using integrated quantities, he found that the same Schmidt law with an index $n=1.4$ holds for both normal and starburst galaxies, covering a range of 5 dex in the gas density and over 6 dex in the SFR; he also found that this relation is mainly driven by the correlation between the SFR and the HI, while there is a poorer correlation between the SFR and the H$_2$, significant only when massive (L$_B>10^{10}$ L$_\\odot$) objects are considered. Owing to the fact that star formation takes place in molecular regions on parsec scales, this result was somewhat unexpected, but Kennicutt argued that the poor knowledge of the CO--to--H$_2$ conversion factor might be responsible for the scatter in the H$_2$/SFR correlation. Using seven galaxies from the BIMA survey \\citep{hel03}, \\citet{won02} studied the applicability of the Schmidt law on local scales, i.e. through a comparison between the SFR and the gas surface density profiles. They found that a Schmidt law between the SFR and the gas content holds also on local scales and, even if the molecular gas content alone does not always control the SFR, they concluded that the correlation found for the total gas is entirely driven by the molecular component. Consistent results were found by \\citet{boi03}, who used low resolution CO data but included an $X$ factor that varies as a function of the metallicity. Another open issue is to asses the influence of the environment on the star formation. Galaxies in clusters suffer from a variety of environmental perturbations, mainly due to gravitational interactions between galaxies themselves or between a galaxy and the cluster potential well \\citep[see the review by][]{bos06} or to various hydrodynamical interactions between the ISM and the intergalactic medium (IGM). Tidal interactions between galaxies or between individual galaxies and the cluster potential \\citep{mer84, byr90} produce both the removal of the outer and looser components and the gas infall towards the galaxy center; harassment \\citep{moo96} can induce sinking of gas towards the galaxy center and can shape the stellar profiles; the ram pressure stripping \\citep{gun72} or the viscous stripping \\citep{nul82} can effectively remove the gas components, mainly from the outer part of the disks. Furthermore, a combination of gravitational and hydrodynamic effects can cause galaxy starvation or strangulation \\citep{lar80,bek02}. According to starvation, the gas that feeds the star formation in the local universe probably comes from the infall of an extended gas reservoir, therefore the effect of removing the outer galaxy halo would be that of preventing further gas infall; as a consequence, the star formation exhausts the available gas, quenching further activity. The literature about the classification and the effects produced by the environment on the evolution of cluster galaxies is very rich \\citep[e.g.][]{bos06}, but, despite many studies on nearby and distant clusters, it is not clear what the dominant processes are, nor how the star formation activity is affected \\citep[and reference therein]{koo04}. Unanimous consent seems only to hold on the statement that environmental processes are effective at removing the outer ISM and therefore affecting the HI component \\citep{cay94,gio85}, leaving the molecular component, well--bounded inside the galaxy potential well, undisturbed \\citep{ken89, bos02c}. In order to study the environmental effects on star formation, we collect both objects that are members of the Virgo cluster and local field galaxies. Since the different components involved in the star formation are observable along a broad stretch of the electromagnetic spectrum, we adopt a multifrequency analysis. To quantify the atomic hydrogen we collect observations at 21 cm; the molecular hydrogen content is estimated indirectly via CO observations at 2 mm; we quantify the stellar mass with observations at the near--infrared (NIR) or visible bands and we study the presence of new stars indirectly, observing the hydrogen recombination line (H$\\alpha$) at 6563 \\AA. The sample is presented in Sec. \\ref{data}, together with the description of the data reduction procedures; the analysis is given in Sec. \\ref{analysis}, while we discuss our results and conclusions in Sec. \\ref{results} and \\ref{concl}. ", "conclusions": "With the aim of exploring the relations between the process of star formation and the physical properties of the ISM in various galaxy environments, we collected state-of-the-art imaging material and maps for 28 massive spiral galaxies belonging to the Virgo cluster and to the local field. The observational material includes: images of the stellar continuum (taken in the red/NIR bands); H$\\alpha$ images of the young ($<4\\times 10^6$ yrs) stars; radio maps at 2mm from the recent {\\it Nobeyama CO atlas of nearby spiral galaxies} that combines good sensitivity to large-scale CO emission with a 15 arcsec spatial resolution, and sensitive radio maps at 21 cm from the ongoing, yet unpublished, VIVA and THINGS surveys carried out at the VLA. Physical parameters have been derived homogeneously and carefully for the individual galaxies applying corrections based on measured quantities, rather than average values. We hope that the effort we put in the analysis of the data reduces the statistical limitations that arise from the paucity of imaging material currently available. \\\\ The present analysis indicates that the bulk of the star formation in spiral galaxies is supported by a diffuse molecular medium which forms through the conversion of the atomic hydrogen, due to the pressure exerted by the stellar potential. However, at the edge of the star forming disks, the HI plays a more important role than previously believed in directly sustaining the star formation activity. When environmental processes cause significant removal of the outer part of the HI disks, galaxies suffer from starvation; the replenishment of the atomic gas inside the optical disks is reduced, leading to a depletion of the molecular component that is consumed during the star formation activity, thus causing the quenching of the star formation activity itself. When the HI removal is so severe that the HI disk shrinks inside the optical radius, there is also a truncation of both the molecular and the star forming disk." }, "0808/0808.0746_arXiv.txt": { "abstract": " ", "introduction": "\\label{ch1:sec1} \\subsection{The Current Paradigm of Early Universe Cosmology} According to the inflationary universe scenario \\cite{Guth} (see also \\cite{Sato,Starob1,Brout}), there was a phase of accelerated expansion of space lasting at least 50 Hubble expansion times during the very early universe. This accelerated expansion of space can explain the overall homogeneity of the universe, it can explain its large size and entropy, and it leads to a decrease in the curvature of space. Most importantly, however, it includes a causal mechanism for generating the small amplitude fluctuations which can be mapped out today via the induced temperature fluctuations of the cosmic microwave background (CMB) and which develop into the observed large-scale structure of the universe \\cite{Mukhanov} (see also \\cite{Press,Sato,Starob2,Lukash}). The accelerated expansion of space stretches fixed co-moving scales beyond the Hubble radius. Thus, it is possible to have a causal mechanism which generates the fluctuations on microscopic sub-Hubble scales. The wavelengths of these inhomogeneities are subsequently inflated to cosmological scales which are super-Hubble until the late universe. The generation mechanism is based on the assumption that the fluctuations start out on microscopic scales at the beginning of the period of inflation in a quantum vacuum state. If the expansion of space is almost exponential, an almost scale-invariant spectrum of cosmological perturbations results, and the squeezing which the fluctuations undergo while they evolve on scales larger than the Hubble radius predicts a characteristic oscillatory pattern in the angular power spectrum of the CMB anisotropies \\cite{Sunyaev}, a pattern which has now been confirmed with great accuracy \\cite{Boomerang,WMAP} (see e.g \\cite{MFB} for a comprehensive review of the theory of cosmological fluctuations, and \\cite{RHBrev2} for an introductory overview). To establish our notation, we write the metric of a homogeneous, isotropic and spatially flat four-dimensional universe in the form \\be \\label{metric} ds^2 \\, = \\, dt^2 - a(t)^2 d{\\bf x}^2 \\, , \\ee where $t$ is physical time, ${\\bf x}$ denote the three co-moving spatial coordinates (points at rest in an expanding space have constant co-moving coordinates), and the scale factor $a(t)$ is proportional to the size of space. The expansion rate $H(t)$ of the universe is given by \\be H(t) \\, = \\, \\frac{{\\dot a}}{a} \\, , \\ee where the overdot represents the derivative with respect to time. \\begin{figure} \\begin{center} \\includegraphics[height=9cm]{canc1.eps} \\caption{Space-time diagram (sketch) of inflationary cosmology. Time increases along the vertical axis. The period of inflation begins at time $t_i$, ends at $t_R$, and is followed by the radiation-dominated phase of standard big bang cosmology. If the expansion of space is exponential, the Hubble radius $H^{-1}$ is constant in physical spatial coordinates (the horizontal axis), whereas it increases linearly in time after $t_R$. The physical length corresponding to a fixed co-moving length scale is labelled by its wave number $k$ and increases exponentially during inflation but increases less fast than the Hubble radius (namely as $t^{1/2}$), after inflation. Hence, the wavelength crosses the Hubble radius twice. It exits the Hubble radius during the inflationary phase at the time $t_i(k)$ and re-enters during the period of standard cosmology at time $t_f(k)$.} \\end{center} \\label{fig:1} \\end{figure} A space-time sketch of inflationary cosmology is shown in Fig. 1. The vertical axis is time. The inflationary phase begins at the time $t_i$ and lasts until the time $t_R$, the time of ``reheating\". At that time, the energy which is driving inflation must change its form into regular matter. The Hubble radius is labelled by $H^{-1}(t)$ and divides scales into those where micro-physics dominates and thus the generation of fluctuations by local physics is possible (sub-Hubble scales) and those where gravity dominates and micro-physical effects are negligible (super-Hubble). As shown in the sketch, during inflation fixed co-moving scales (labelled by $k$ in the sketch) are inflated from microscopic to cosmological. Note also that the horizon, the forward light cone, becomes exponentially larger than the Hubble radius during the inflationary phase. \\subsection{Challenges for String Cosmology} Working in the context of General Relativity as the theory of space-time, inflationary cosmology requires the presence of a new form of matter with a sufficiently negative pressure $p$ ($p < - 2/3 \\rho$, where $\\rho$ denotes the energy density). In order to obtain such an equation of state, in general the presence of scalar field matter must be assumed. In addition, it must be assumed that the scalar field potential energy dominates over the scalar field spatial gradient and kinetic energies for a sufficiently long time period. The Higgs field used for the spontaneous breaking of gauge symmetries in particle physics has a potential which is not flat enough to sustain inflation. Models beyond the Standard Model of particle physics, in particular those based on supersymmetry, typically have many scalar fields. Nevertheless, it has proven to be very difficult to construct viable inflationary models. The problems which arise when trying to embed inflation into the context of effective field theories stemming from superstring theory are detailed in the contribution to this book by Burgess. If inflationary cosmology is realized in the context of classical General Relativity coupled to scalar field matter, then an initial cosmological singularity is unavoidable \\cite{Borde}. Resolving this initial singularity is one of the challenges for string cosmology. The energy scale during inflation is set by the observed amplitude of the CMB fluctuations. In simple single field models of inflation, the energy scale is of the order of the scale of Grand Unification, i.e. many orders of magnitude larger than scales for which field theory has been tested experimentally, and rather close to the string and Planck scales, scales where we know that the low energy effective field theory approach will break down. It is therefore a serious concern whether the inflationary scenario is robust towards the inclusion of non-perturbative stringy effects, effects which we know must not only be present but in fact must dominate at energy scales close to the string scale. The problem for cosmological fluctuations is even more acute: provided that the inflationary phase lasts for more than about 70 Hubble expansion times, then all scales which are currently probed in cosmological observations had a wavelength smaller than the Planck length at the beginning of the inflationary phase. Thus, the modes definitely are effected by trans-Planckian physics during the initial stages of their evolution. The ``trans-Planckian problem\" for fluctuations \\cite{RHBrev1,Jerome} is whether the stringy effects which dominate the evolution in the initial stages leave a detectable imprint on the spectrum of fluctuations. To answer this question one must keep in mind that the expansion of space does not wash out specific stringy signatures, but simply red-shifts wavelengths. For string theorists, the above ``trans-Planckian problem\" is in fact a window of opportunity: if the universe underwent a period of inflation, this period will provide a microscope with which string-scale physics can be probed in current cosmological observations. Some of the conceptual problems of inflationary cosmology are highlighted in Figure 2, a space-time sketch analogous to that of Figure 1, but with the two zones of ignorance (length scales smaller than the Planck (or string) length and densities higher than the Planck (or string) density) are shown. As the string scale decreases relative to the Planck scale, the horizontal line which indicates the boundary of the super-string density zone of ignorance approaches the constant time line corresponding to the onset of inflation. This implies that the inflationary background dynamics itself might not be robust against stringy corrections in the dynamical equations. The sketch in Figure 2 also shows the exponential increase of the horizon compared to the Hubble radius during the period of inflation. \\begin{figure} \\begin{center} \\includegraphics[height=9cm]{infl3.eps} \\caption{Space-time diagram (sketch) of inflationary cosmology including the two zones of ignorance - sub-Planckian wavelengths and trans-Planckian densities. The symbols have the same meaning as in Figure 1. Note, specifically, that - as long as the period of inflation lasts a couple of e-foldings longer than the minimal value required for inflation to address the problems of standard big bang cosmology - all wavelengths of cosmological interest to us today start out at the beginning of the period of inflation with a wavelength which is in the zone of ignorance.} \\end{center} \\label{fig:2} \\end{figure} \\subsection{Preview} The conceptual problems of inflationary cosmology discussed in the previous subsection motivate a search for a new paradigm of early universe cosmology based on string theory. Such a new paradigm may provide the initial conditions for a robust inflationary phase. However, it may also lead to an alternative scenario. In the following, we will explore this second possibility. In the best possible world, the initial phase of string cosmology will eliminate the cosmological ``Big Bang\" singularity, it will provide a unified description of space, time and matter, and it will allow a controlled computation of the induced cosmological perturbations. The development of such a consistent framework of string cosmology will, however, have to be based on a consistent understanding of non-perturbative string theory. Such an understanding is at the present time not available. Given the lack of such an understanding, most approaches to string cosmology are based on treating matter using an effective field theory description motivated by string theory. However, in such approaches key features of string theory which are not present in field theory cannot be seen. The approach to string cosmology discussed below is, in contrast, based on studying effects of new degrees of freedom and new symmetries which are key ingredients to string theory, which will be present in any non-perturbative formulation of string theory. ", "conclusions": "The string gas scenario is an approach to early universe cosmology based on coupling a gas of strings to a classical background. It includes string degrees of freedom and string symmetries which are hard to implement in an effective field theory approach. The background of string gas cosmology is non-singular. The temperature never exceeds the limiting Hagedorn temperature. If we start the evolution as a dense gas of strings in a space in which all dimensions are string-scale tori, then there are dynamical arguments according to which only three of the spatial dimensions can become large \\cite{BV}. Thus, string gas cosmology yields the hope of understanding why - in the context of a theory with more than three spatial dimensions - exactly three are large and visible to us. If the Hagedorn phase (the phase during which the temperature is close to the Hagedorn temperature and both the scale factor and the dilaton are static) is sufficiently long to establish thermal equilibrium on length scales of about 1 mm, then string gas cosmology can provide an alternative to cosmological inflation for explaining the origin of an almost scale-invariant spectrum of cosmological fluctuations \\cite{NBV}. A distinctive signature of the scenario is the slight blue tilt in the spectrum of gravitational waves which is predicted \\cite{BNPV1}. The inflationary universe scenario has successes beyond the fact that it successfully predicted a scale-invariant spectrum of fluctuations - it also explains why, starting from a hot Planck scale space, an extremely low entropy state - one can obtain a universe which is large enough and contains enough entropy to correspond to our observed universe. In addition, it explains the observed spatial flatness. However, if the Hagedorn phase of string gas cosmology is realized as a long bounce phase in a universe which starts out large and cold, then the horizon, flatness, size and entropy problems do not arise. A serious concern for the current realization of string gas cosmology, however, is the gravitational Jeans instability problem. This problem was first raised in \\cite{stability}. In the context of dilaton gravity, it can be shown \\cite{dilflucts} that gravitational fluctuations do not grow. However, dilaton gravity is not a consistent background for the Hagedorn phase of string gas cosmology. One might hope that since the string states are relativistic, the gravitational Jeans length will be comparable to the Hubble radius, as it is for a gas of regular radiation. However, a recent computation of the speed of sound in string gas cosmology \\cite{Nima} has shown that in a background space sufficiently large to evolve into our present universe the overall speed of sound is very small. Further work needs to be done on this issue. This is complicated by fact that string thermodynamics is non-extensive (see e.g. \\cite{Cobas2}), which leads to problems in using the usual thermodynamic intuition. Note that the background space does not need to be toroidal. Crucial for string gas cosmology to yield the predictions summarized above is the existence and stability (or quasi-stability) of string winding modes. Certain orbifolds \\cite{Col1} have also been shown to yield good backgrounds for string gas cosmology. Non-trivial one cycles will ensure the existence and stability of string winding modes. If the background space does not have have any non-trivial one-cycles, then it might be possible to construct a cosmological scenario based on stable branes rather than strings. The cosmology of brane gases has been considered in \\cite{branes}. If there are stable winding strings, and if the string coupling constant is small such that the fundamental strings are lighter than branes, then \\cite{ABE} it is the fundamental strings which will dominate the thermodynamics in the Hagedorn phase and which will be the most important degrees of freedom for cosmology. However, if there are no stable winding strings, then winding branes would become important. It appears at the present time that Heterotic string theory is most suited for string gas cosmology since this theory admits the enhanced symmetry states which have been shown to yield a very simple way to stabilize the size and shape moduli of the extra spatial dimensions. It will be interesting to study if string gas cosmology can be embedded into particular models of Heterotic string theory which yield reasonable particle phenomenology. The presentation we gave of string gas cosmology is based on minimal input. In particular, we did not include fluxes since we assume that the net fluxes should cancel for a situation with the most symmetric initial conditions. The role of fluxes in string gas cosmology has been studied in \\cite{Campos}. Whereas the primary application of string gas cosmology will be to the cosmology of the very early universe, it is also interesting to consider applications of string gas cosmology to later time cosmology. The late time dynamics of string and brane gases has been considered in \\cite{late}. In particular, in \\cite{Ferrer} applications of string gases to the dark energy problem has been considered (see also \\cite{McInnes}). String and brane gases have also been studied as a way to obtain inflation \\cite{Turok,Parry,Anupam} (see also \\cite{Freese}), or as a way to obtain non-inflationary bulk expansion which may provide a way to solve the size problem in string gas cosmology if one starts with a spatial manifold of string scale in all directions \\cite{Natalia}. \\vskip0.3cm \\centerline{\\bf Acknowledgements} \\vskip0.2cm I am grateful to all of my present and former collaborators with whom I have had the pleasure of working on string gas cosmology. For comments on the draft of this manuscript I wish to thank Nima Lashkari and Subodh Patil. This work is supported in part by an NSERC Discovery Grant and by funds from the Canada Research Chairs Program." }, "0808/0808.1740_arXiv.txt": { "abstract": "The broad-line radio galaxy 3C\\,111 has been suggested as the counterpart of the $\\-$--ray source 3EG\\,J0416+3650. While 3C\\,111 meets most of the criteria for a high-probability identification, like a bright flat-spectrum radio core and a blazar-like broadband SED, in the Third EGRET Catalog, the large positional offset of about 1.5$^\\circ$ put 3C\\,111 outside the 99\\% probability region for 3EG\\,J0416+3650, making this association questionable. We present a re-analysis of all available data for 3C\\,111 from the EGRET archives, resulting in probable detection of high-energy $\\gamma$--ray emission above 1000\\,MeV from a position close to the nominal position of 3C\\,111, in two separate viewing periods (VPs), at a 3$\\sigma$ level in each. A new source, GRO\\,J0426+3747, appears to be present nearby, seen only in the $>$1000\\,MeV data. For $>$100\\,MeV, the data are in agreement with only one source (at the original catalog position) accounting for most of the EGRET-detected emission of 3EG\\,J0416+3650. A follow-up \\textit{Swift} UVOT/XRT observation reveals one moderately bright X--ray source in the error box of 3EG\\,J0416+3650, but because of the large EGRET position uncertainty, it is not certain that the X--ray and $\\gamma$--ray sources are associated. A \\textit{Swift} observation of GRO\\,J0426+3747 detected no X--ray source nearby. ", "introduction": "One of the main scientific goals of the recently-launched $\\gamma$--ray astronomy satellite mission GLAST and its Large Area Telescope (LAT) \\citep{Ri07, Mi07} is to shed light on the nature of powerful relativistic extragalactic jets, which are ejected from the nuclei of some active galaxies (AGN). Originally, this class of AGN was defined based on the bright and prominent radio emission from these jets: the radio-loud population of active galaxies. Based on data from EGRET, the high-energy $\\gamma$--ray telescope on the {\\it Compton Gamma Ray Observatory}, it was realized \\citep{Fi94, Th94, vM95} that the largest population of extragalactic $\\gamma$--ray sources in the GeV regime is represented by those radio-loud AGN whose jets are pointed at a small angle to the line of sight. An intimate link is tied between radio VLBI and high-energy $\\gamma$--ray astronomy by the fact that the bright compact radio emission of these so-called blazars provides excellent targets for parsec-scale resolution VLBI observations of their jet structure. Of special interest are the questions of where in the AGN jets the bright $\\gamma$--ray emission is produced, and how the emission is interacting with its immediate environment and with other parts of the jet. This knowledge would enable us to put crucial constraints on the processes of jet formation, collimation, and acceleration. GLAST is expected to yield densely sampled $\\gamma$--ray light curves of hundreds of extragalactic jets that are bright enough to be detected on time scales of days to weeks, and thousands on time scales of months to years. Most of these objects will be blazars; in fact, all but two of the firmly identified extragalactic EGRET sources (LMC, Sreekumar et al. 1992; Centaurus\\,A, Sreekumar et al. 1999) are blazars. In contrast to the blazars, most radio galaxies have larger inclination angles. That allows better (deprojected) linear resolution with VLBI observations, and in the case of stratified jet structures, it allows observations of the slower jet layers, e.g. a sheath, whose emission may be swamped by the much brighter beamed emission from faster jet regions, e.g. a fast spine, in blazars. \\citet{Ghi05} have presented such a spine-sheath stratified-jet model, predicting detectable $\\gamma$--ray emission from the nuclei of some radio galaxies. In their model, each of the two emission regions sees photons coming from the other part relativistically enhanced because of the relative speed difference, giving rise to an additional inverse-Compton emission component. In this letter we provide additional support for the identification of the broad-line radio galaxy 3C\\,111 as a $\\gamma$--ray source, responsible for a portion of the source 3EG\\,J0416+3650. Such an association was suggested as possible in the third EGRET catalog \\citep[3EG]{Ha99}, but was considered unlikely because of the large positional offset of 3C\\,111 from 3EG\\,J0416+3650. The present work was stimulated by the recent report of \\citet{Sg05}, who used multiwavelength data to strengthen the case for the association between 3C\\,111 and 3EG\\,J0416+3650. We show here that 3EG\\,J0416+3650 is most likely composed of at least two, and more likely three, separate sources, one of which is in good positional agreement with 3C\\,111. As demonstrated recently, based on VLBA monitoring data at $\\lambda 2$\\,cm from the MOJAVE program \\citep{Kad08}, the parsec-scale jet of 3C\\,111 shows a variety of physically different regions in a relativistic extragalactic jet, such as a compact core, superluminal jet components, recollimation shocks, and regions of interaction between the jet and its surrounding medium, which are all possible sites of $\\gamma$--ray production. Its relatively large inclination angle of $\\sim$19$^\\circ$ makes 3C\\,111 a particularly well-suited target for tests of structured-jet models, such as the model of \\citet{Ghi05}. We describe our re-analysis of the available EGRET data on 3C\\,111, as well as follow-up \\textit{Swift} OVOT and XRT observation, in Sect.~\\ref{sect:analysis}, and discuss our results and their implications in Sect.~\\ref{sect:conclusions}. ", "conclusions": "\\label{sect:conclusions} The $\\gamma$--ray source 3EG\\,J0416+3650 seems to be composed of at least three variable sources. One of the sources, detected only above 1000\\,MeV, is close to the radio galaxy 3C\\,111, and plausibly associated with it; a new source, GRO\\,J0426+3747, is also seen only above 1000\\,MeV, and has no obvious identification at other wavelengths. Most of the flux $>$100\\,MeV is from the third source, for which there is no evidence in the $>$1000\\,MeV data. The position obtained here for the third source is very near the catalog position of 3EG\\,J0416+3650. It might be associated with the XRT source Swift\\,J041554.3+364926, but the EGRET position uncertainty makes a firm association impossible. The region under study is $8-10^\\circ$ from the Galactic plane, so either 3EG\\,J0416+3650 or GRO\\,J0426+3747 (or both) could be local. A recent stacking search found no evidence for $\\gamma$--ray emmision from radio galaxies as a class \\citep{Cil04}. That paper specifically excluded 3C\\,111 from consideration because of its blazar-like superluminal motion, so there is no direct conflict with the results presented here. That study made no systematic cut on the angle to the line-of-sight of the radio jet, so for most of the galaxies included, that angle was considerably larger than the $\\sim$19$^\\circ$ which has been estimated for 3C\\,111 \\citep{Kad08}. Since a larger line-of-sight angle implies substantially smaller Doppler boosting and therefore lower $\\gamma$--ray output, there is no obvious conflict between the present results and those of \\citet{Cil04}. A more significant challenge is posed by the lack of EGRET detection of $\\gamma$--ray emission from the (much closer) AGN M87. Superluminal motion has been detected in the jet knot HST-1, about 0.86\\,arcsec downstream from the radio core, both in the optical and radio bands \\citep{Bir99,Che07}. Most estimates of our line-of-sight angle to its jet are in the range $30-45^\\circ$. If we assume that M87 is similar to 3C\\,111, it is not clear whether its Doppler boosting is sufficiently low to compensate for its much smaller distance, and thereby account for its non-detection by EGRET. Our re-analysis of the available EGRET data supports the previous association of the source 3EG\\,J0416+3650 with the broad line radio galaxy 3C\\,111, but with 3C\\,111 responsible for only a portion of the $\\gamma$--ray emission. Furthermore, it explains the relatively large positional offset noted in \\citet{Ha99}. We have compiled an historical SED of 3C\\,111 which shows that the X--ray data may well extrapolate into the EGRET range, particularly during flares, which is in agreement with the intermittent nature of detections in the individual EGRET viewing periods. It is interesting to note that we detect 3C\\,111 at almost exactly the $\\gamma$--ray flux that is predicted by equation (12) in \\citet{Ghi05}: $(1.41-14.1) \\times 10^{-11}$\\,erg\\,s$^{-1}$\\,cm$^{-2}$, scaling from the 5\\,GHz values in Table~\\ref{tab:3c111_sed}. Note that 3C\\,111 (z=0.0485) would be the most distant radio galaxy detected in $\\gamma$ rays, about twice as far as NGC\\,6251 (Mukherjee et al. 2002; z=0.0247). The detection of $\\gamma$--ray emission from 3C\\,111 further supports the hypothesis that radio galaxies may represent an important class of LAT sources. GLAST was launched on 11 June, 2008; its LAT detector will continuously scan the entire sky over a 3-hr interval. With a factor of 30 greater sensitivity than EGRET, and with a factor of $\\sim$3 better PSF above 1\\,GeV, it will be more efficient than EGRET for detecting transients over the whole sky. In the case of 3C\\,111, GRO\\,J0426+3747, and 3EG\\,J0416+3650, the smaller PSF and greater effective area of the LAT will clearly separate these three sources." }, "0808/0808.1789_arXiv.txt": { "abstract": "We present U, B, V, R, I, H$\\alpha$ and NUV photometry of 14 galaxies in the very local Universe (within 10 Mpc). Most objects are dwarf irregular galaxies (dIrr) and are probably associated with the NGC 672/IC 1727 and NGC 784 galaxy groups. The galaxies are at low redshift (51$\\leq$v$_{\\odot} \\leq$610 km sec$^{-1}$) and most appear projected on the sky as a six degree long linear filament. We show that the galaxy positions along this filament correlate with their radial velocity, hinting to an interpretation as a single kinematic entity. Our CCD photometry indicates that all objects qualify as ``dwarf galaxies'' with M$_B\\geq-18$ mag. We examine the star formation (SF) properties of individual objects in the context of their immediate environment. The current SF rate (SFR) is derived directly from the H$\\alpha$ line flux. An approximate SF history is derived by comparing the multi-band photometry with results of galaxy evolution models from Bruzual \\& Charlot (2003a, 2003b), assuming short SF bursts separated by long quiescence periods. Relations between the current SFR and the HI mass or the absolute B magnitude for the galaxies in these groups indicate that these objects behave like normal galaxies. A comparison of the photometric measurements with evolutionary synthesis model predictions indicates that most objects can be understood as containing at least one ``old'' stellar population ($\\geq$1-10 Gyr) and one ``young'' population ($\\leq$30 Myr). For both groups, the recent SF bursts appear to have occurred at similar times, a few to a few 10s of Myr ago, arguing for synchronicity in star formation in these objects. In an attempt to evaluate the possible role of galaxy-galaxy interaction, we investigate the trend of the SFR with an object's projected distance from the brightest and most massive galaxies of each group. We do not find a steadily decreasing star formation as function of this distance; such a result could be expected if the star formation would have been triggered by interactions. We propose that one possible explanation of the $\\sim$synchronous star formation in all objects is accretion of cold gas from intergalactic space onto dark matter haloes arranged along a filament threading the void where these dwarf galaxies reside. We point out this galaxy sample as an ideal target to study hierarchical clustering and galaxy formation among very nearby objects. ", "introduction": "Star formation is one of the more important processes in the Universe. Studying both the SF history in galaxies and the influence of the environment on this process contributes to the understanding of cosmic evolution. In the past few decades increasing efforts have been made to better understand the dependence of SF on local and environmental properties. Tidal interactions were promoted as triggers and enhancers of SF (e.g., Larson \\& Tinsley 1978, Li et al.\\ 2008). In practice, the influence of tidal interactions is more complex, depends upon various factors, and varies among different environments. Hashimoto et al.\\ (1998) studied the variation and dependence of these influences in different environments for different types of galaxies and found different SFRs in the field and in clusters, even when the galaxy densities are similar. They also found that for a given galaxy concentration index, galaxies in lower-density environments show higher SFR levels than objects in higher galaxy density regions. At least two processes were found to influence the susceptibility of SF to the environment: gas-removal processes responsible for the variation with galaxy density of the SFRs of normal galaxies, and galaxy-galaxy interactions responsible for the prevalence of SF bursts in intermediate-density environments. At low redshifts, galaxies in groups may show a reverse relation with a somewhat higher SFR (but not extreme burst conditions) the more isolated a galaxy is (see Martig \\& Bournaud 2008). However, Brosch et al.\\ (2004) examined specifically the influence of interactions on SF in dwarf galaxies and concluded that galaxy interactions are probably not a primary trigger of present SF in such objects. About half the galaxies in the Universe are probably in groups (Mamon 2007) and these constitute an interesting environment for the SF investigation. The evolution of galaxies within groups may depend upon the density of the group and the level of interaction between its galaxies. Many groups consist of normal galaxies surrounded by clouds of dwarf galaxies (Miller 1996, Cote et al.\\ 1997, Mateo 1998). The dwarf galaxies (DGs) may have formed due to tidal interaction of individual galaxies or of groups of galaxies, or as leftover material from the colliding and merging of more massive galaxies (e.g., Hunsberger et al.\\ 1996, Bournaud \\& Duc 2006). They may also represent early stages of hierarchical clustering, with the stars that provide the observed luminosity forming in a low-mass halo. The DGs in a group contain usually only a few percents of the group mass and in many cases may be accreted by larger galaxies. Although dwarf galaxies constitute the majority of the galaxy population, large uncertainties surround their formation and evolutionary histories. The dwarf galaxy population follows a strong morphology-density relation, with passively evolving systems mostly found in close proximity to massive galaxies, in contrast to the more widespread gas-rich, star forming population. Studying galaxy groups, in particular the dwarf-rich ones, contributes to the understanding of cosmic structure and evolution, since such groups constrain cosmological models, imply the shape of the dark halos around the massive galaxies, and play a role in the evolution of these galaxies (see Bournaud \\& Duc 2006 and references therein). The present study complements similar ones on other groups of galaxies in the local Universe, such as for the NGC 628 and M81 groups (e.g., Sharina et al.\\ 2006, Karachentsev \\& Kaisin 2007). Hickson (1982) published a catalog of compact groups of galaxies in the local Universe, and investigated various characteristics of these groups (e.g., Hickson et al.\\ 1992, de Oliveira \\& Hickson 1991, 1994). Samples of galaxies in other environments have been examined, such as in the Virgo cluster (e.g., on dwarf galaxies, Almoznino \\& Brosch 1998a, b; Brosch, Heller \\& Almoznino 1998; Heller, Almoznino \\& Brosch 1999; Heller \\& Brosch 2001). The influences of galaxy interactions on star formation were reviewed by Larson \\& Tinsley (1978), Hashimoto et al.\\ (1998) and Li et al.\\ (2008). In most previous papers the galaxies were either normal (i.e., large spirals, lenticulars or ellipticals) or dwarf galaxies; the specific aspect the present paper adopts is to study these properties in a fairly isolated group or groups of DGs. Tidal interactions between two strongly-interacting major galaxies tend to form sometimes rather concentrated clouds of DG satellites around them, or long tails of DGs as seen e.g., in the Hickson compact group 100 (de Mello et al. 2008). If one would observe dwarf galaxies in a group apparently aligned on the sky in a linear configuration, and with no major galaxies in the immediate neighborhood, this could imply the presence of a dark matter filament onto which the intergalactic matter collapses, forming the observed DGs. This work studies a sample galaxies in the very local Universe ($\\sim$10 Mpc) consisting of the NGC 672/IC 1727 group of galaxies, the NGC 784 group of galaxies, and several other galaxies in the same vicinity with similar redshifts. Our attention was drawn to this sample when inspecting the precursor observations of the ALFALFA survey (Giovanelli et al. 2005) because of its obvious linear structure that could qualify as a ``filament'' of galaxies, and because of its apparent isolation. The galaxies are located approximately in the anti-Virgo direction, in a region of the Universe that is a void in the galaxy distribution; this is the ``Local Mini-void'' (Karachentsev et al. 2004). As will be shown below, all galaxies are ``dwarfs'' (M$_B\\geq$-18 mag) and most are dIrr (e.g., Karachentseva \\& Karachentsev 1998, Huchtmeier et al. 2000). We study the SF properties and SF history of each galaxy using surface photometry, and examine general similarities and dependencies on the local, environmental and mutual properties. Current $\\Lambda$CDM simulations predict that low-amplitude filamentary structures criss-cross the voids (Peebles 2007). The galaxies corresponding to the halos in those filaments are expected to be low-luminosity, star-forming galaxies (Hoyle et al. 2005). Saintonge et al. (2008) analyzed the ALFALFA catalog covering a portion of the nearby void in front of the Pisces-Perseus Supercluster at cz$\\sim$2000 km sec$^{-1}$ and, within a volume of 460 Mpc$^3$, did not detect a single galaxy. In contrast, one could expect to detect 38 HI sources in such a volume based on scaling the predictions of Gottl\\\"{o}ber et al. (2003) with a dark-to-HI mass ratio of 10:1. Dekel \\& Birnboim (2006) discussed the bimodality observed in galaxy properties about a characteristic stellar mass of $3\\times10^{10}$ M$_{\\odot}$. The bimodality implies that less massive galaxies tend to be ungrouped, blue, star-forming disks and would be formed in low galaxy-density regions, while more massive galaxies are typically grouped, red, old-star spheroids and would reside in clusters and in denser groups. In haloes below a critical shock-heating mass of 10$^{12}$ M$_{\\odot}$ disks are built by cold-gas streams, not heated by a virial shock, yielding efficient and prompt star formation. The Dekel \\& Birnboim scenario explains naturally why one could expect to see dwarf galaxies even in regions that are not conducive to intense tidal interactions, provided dark matter haloes of the right mass are present and HI material for SF is available. In what follows we argue that the objects studied here may be the first case in the local Universe where we witness the first stages of aggregation of matter that will, eventually, form a major galaxy. The plan of the paper is as follows: in Section~\\ref{sec.sample} we describe the galaxy sample, the observations, mostly collected at the Wise Observatory, and the data reduction, in \\S~\\ref{sec.results} we present the results and discuss individual objects, in \\S~\\ref{sec.discuss} we discuss the results and present arguments in favour of the interpretation proposed above, and in \\S~\\ref{sec.summary} we conclude and summarize our findings. \\section {Observations and data reduction} \\label{sec.sample} \\subsection {The sample} The sample consists of the catalogued NGC 672/IC 1727 and NGC 784 galaxy groups and of several additional galaxies in the same sky area and at similar redshifts; these are listed in Table \\ref{gal}. The galaxies were chosen primarily from the ALFALFA survey (see Giovanelli et al.\\ 2005, Saintonge et al.\\ 2008). % ALFALFA is an unbiased HI survey of the extragalactic sky visible from Arecibo. The precursor ALFALFA observations (Giovanelli et al.\\ 2005) covered the sky area (0h$\\leq$RA$\\leq$6h) and (26$^{\\circ} \\leq \\delta\\leq 28.5^{\\circ}$) and 10 galaxies in the vicinity of NGC 672/IC 1727 with similar redshifts (cz$\\leqslant$600 km sec$^{-1}$) were included among the 166 identified objects. Figure \\ref{fig:galaxies} shows the projection of the different objects on the sky, with the redshift of each object indicated. It is clear that the galaxies form some kind of linear structure, which we call ``filament''; the projected distribution is approximately linear with the objects at the north-east side showing generally lower redshifts. The length of the filament is $\\sim$six degrees; a few galaxies diverge significantly from the linear distribution and we consider them to be ``isolated''. The distance to the galaxies is of order 5 Mpc; the $\\sim6^{\\circ}$ projected angular extent translates into a projected physical length for the linear structure of $\\sim$500 kpc. % As already mentioned, most of the galaxies studied here were previously classified as dwarf irregulars (e.g., Karachentseva \\& Karachentsev 1998, Huchtmeier et al.\\ 2000). In Table~\\ref{gal} we list not only the galaxy name, which in some cases is taken from the Arecibo Galaxy Catalogue (AGC: a private compilation of Haynes and Giovanelli maintained at Cornell), but also the heliocentric velocity in km sec$^{-1}$ and the 50\\% width of the HI line (also in km sec$^{-1}$) taken from Saintonge et al. (2008) or from the HyperLEDA data base (Paturel et al. 2003). Column 7 lists the position angle (PA) of the major axes of the objects, determined from a visual inspection of the blue galaxy images on-line at the Canadian Astronomy Data Centre (CADC). The PA is measured in degrees, clockwise from North through West. AGC 111945 was too faint and small to derive its PA. The last column of the table lists the galaxy group to which an object could belong. \\begin{table}[!h] \\caption{The sample galaxies arranged by group membership} \\vspace{0.5cm} \\label{gal} \\begin{footnotesize} \\begin{center} \\begin{tabular}{|c|c|c|c|c|c|c|c|} \\hline Galaxy & ALFALFA name & $\\alpha$[hh:mm:ss.s] & $\\delta$[$^\\circ$:':''] & v$_{\\odot}$ & w$_{50}$ & PA ($^\\circ$) & Group\\\\ \\hline NGC 672 & HI014753.9+272555 & 01:47:54.5 & +27:25:58 & 429 & 205 & 332 & NGC 672\\\\ IC 1727 & HI014729.9+271958 & 01:47:29.9 & +27:20:00 & 330 & 115 & 35 & NGC 672\\\\ AGC 110482 & HI014214.9+262202 & 01:42:17.3 & +26:22:00 & 357 & 30 & 53 & NGC 672\\\\ AGC 111945 & HI014441.4+271707 & 01:44:42.7 & +27:17:18 & 423 & 36 & ? & NGC 672\\\\ NGC 111946 & HI014640.9+264754 & 01:46:42.2 & +26:48:05 & 367 & 21 & 0 & NGC 672\\\\ AGC 112521 & HI014105.8+272007 & 01:41:08.0 & +27:19:20 & 274 & 26 & 291 & NGC 672\\\\ LEDA169957 & --- & 01:36:35.9 & +23:48:54 & 563 & 47 & 307 & Isolated\\\\ NGC 784 & HI020115.0+284953 & 02:01:16.9 & +28:50:14 & 193 & 88 & 0 & NGC 784\\\\ AGC 111977 & HI015519.2+275645 & 01:55:20.4 & +27:57:13 & 207 & 29 & 297 & NGC 784\\\\ AGC 111164 & HI020009.3+284954 & 02:00:10.2 & +28:49:53 & 164 & 27 & 24 & NGC 784\\\\ UGC 1281 & --- & 01:49:32.0 & +32:35:23 & 156 & 98 & 306 & NGC 784\\\\ AGC 122834 & HI020347.0+291153 & 02:03:47.0 & +29:11:53 & 51 & 10 & 297 & NGC 784?\\\\ AGC 122835 & HI020533.0+291358 & 02:05:33.0 & +29:13:58 & 50 & 23 & - & N784; no Opt ID \\\\ UGC 1561 & --- &02:04:05.1 & +24:12:30 & 610 & 47 & 90 & Isolated\\\\ NGC 855 & HI021404.3+275302 & 02:14:03.6 & +27:52:36 & 594 & 81 & 305 & Isolated\\\\ \\hline \\end{tabular} \\end{center} \\end{footnotesize} \\end{table} \\begin{figure}[t] \\centering{ \\includegraphics[width=12cm]{Gal3DwithVel.eps} \\caption{The galaxies as projected on the sky, represented by circles of different sizes. The circle sizes are proportional to the galaxies' semi-major axes. Redshifts are marked next to each galaxy symbol. A linear structure is evident from the distribution of 11 of the 14 objects plotted here.} \\label{fig:galaxies}} \\end{figure} Though we found no previous results covering the SF in the sample galaxies, some objects were investigated before: the NGC 672/IC 1727 pair and their interaction have been studied by de Vaucouleurs et al.\\ (1976), Combes et al.\\ (1980), Hodge \\& Kennicutt (1983), Sohn \\& Davidge (1996), Karachentseva \\& Karachentsev (1998); NGC 784 and some members of its group have been studied by Karachentseva \\& Karachentsev (1998), Huchtmeier et al.\\ (2000), Tikhonov \\& Karachentsev (2006), Tully et al.\\ (2006); other galaxies of the sample were studied by e.g., Wallington et al.\\ (1988), Huchtmeier \\& Richter (1989), Moustakas \\& Kennicutt (2006). Most galaxies were measured in HI by the ALFALFA survey (e.g., Giovanelli et al.\\ 2005) or are included in Karachentsev's catalog of neighboring galaxies (Karachentsev et al. 2004). Saintonge et al.\\ (2008) studied the HI properties of the NGC 672/IC 1727 and NGC 784 galaxy groups among other objects. Saintonge et al.\\ (2008) included the NGC 672/IC 1727 and the NGC 784 groups along with two new candidate members of the NGC 784 group among the objects they studied. Only one of those two candidates, AGC 122834, has an optical counterpart. The other, AGC 122835 at 02:05:33.0 +29:13:58, is also listed in Table~\\ref{gal}. In addition to the 12 objects retrieved from the ALFALFA data sets, NED was queried for objects up to 300 arcmin away and with similar velocities; this added three additional neighboring galaxies. All the objects fall within the ``Local Mini-void'' Karachentsev et al.\\ (2004, Fig. 4). Note that the ALFALFA survey is likely to find even more members of this filament once more survey data for other declination strips will be released and redshifts of low surface brightness objects will be measured. The literature contains a number of distance estimates to objects in our sample. Sohn \\& Davidge (1996) assigned a distance of 7.9$^{+1.0}_{-0.9}$ Mpc to NGC 672 based on the brightness of the red supergiants. Karachentsev et al.\\ (2004) reported 7.2 Mpc to the NGC 672/IC 1727 pair based on the luminosity of the brightest stars. A distance of 5.0 Mpc was measured in a similar way for NGC 784 by Drozdovsky \\& Karachentsev (2000). The distance to AGC 111977 and AGC 111164 was reported as 4.7 Mpc by Karachentsev et al.\\ (2004) using the red giant branch method. The distances to UGC 1281 was estimated by Karachentsev et al.\\ (2004) to be 5.4 Mpc using the brightest stars. The reported distance for NGC 855 (using the surface brightness fluctuations method) was 9.73$\\pm$1.7 Mpc (Tonry et al. 2001). The question of distances to these galaxies was discussed by Giovanelli et al. (2005) who used the peculiar velocity model of the local Universe of Tonry et al. (2000). The model provides the expected peculiar velocity at a given point in space, but can be inverted to yield the distance to a galaxy at certain celestial coordinates and with a measured redshift. Giovanelli al. caution that the ``thermal'' rms peculiar velocity of galaxies in the Tonry et al. model translates into a distance uncertainty of about 2.7 Mpc and thus the distances to nearby objects, and in particular to the NGC 672 and NGC 784 groups, are highly uncertain. Since the distance estimates were produced by a variety of methods, one wonders whether the distance differences between members of the sample is believable. In particular, when Karachentsev et al.\\ (2004) quote 4.7 Mpc to AGC 111977 and AGC 111164 and 9.73 Mpc to NGC 855, can we indeed assume that these objects are separated by $\\sim$5 Mpc in depth? Is it possible that the objects could be at about the same distance, despite their somewhat different redshifts? Given the uncertainty in distance, we decided to disregard peculiar velocities and use distances assuming only Hubble expansion with H$_{0}$=73 km sec$^{-1}$/Mpc to evaluate intrinsic parameters for each galaxy. The HI masses and the SFRs were calculated assuming these distances. However, in the discussion of the 3D structure traced by the galaxies we retain the option of considering the objects to be aligned along a filament (that may be inclined to the line of sight) and being at approximately the same distance from us. \\begin{figure}[t] \\centering{ \\includegraphics[width=12cm]{VelocityVsX.eps} \\caption{Position-velocity diagram for all the galaxies as projected on the sky. The horizontal axis marked X is the distance along a line fitted to the galaxy positions. The two symbols at the top-right part of the diagram are UGC 1561 and NGC 855.} \\label{fig:gals_X_vs_Vel}} \\end{figure} We plot in Figure~\\ref{fig:gals_X_vs_Vel} the position of each galaxy along the line fitted to their projected location on the sky vs. the heliocentric radial velocity of each object. The plot shows that most galaxies line up nicely on a $\\sim$6$^{\\circ}$ linear feature, with those at higher Right Ascension having generally a lower radial velocity. This kinematic relation can be understood as additional evidence that the objects are probably part of the same large structure in the nearby Universe. The two points at the upper right in Figure~\\ref{fig:gals_X_vs_Vel} are UGC 1561 and NGC 855; it is possible that these might not be part of the same kinematic structure delineated by the other galaxies. \\subsection {Observations} All broad-band and H$\\alpha$ observations were performed with the Wise Observatory (WiseObs) 1-m telescope between September 2006 and February 2008. We obtained CCD images of all 14 galaxies of the sample using the PI VersArray camera with a $1340\\times1300$ pixel thinned and back-illuminated CCD and a scale of 0.58\"/pixel at the f/7 Ritchey-Chr\\'{e}tien focus. The CCD saturation level is 65535 counts per pixel, the gain is 2.1 $e^{-}$/ADU, and the readout noise is 2.87 $e^{-}$ per pixel. The seeing on most nights was 2.2\"-2.8\". Images of each galaxy were obtained through UBVRI and zero-redshift H$\\alpha$ filters. Landolt (1992) and spectrophotometric (e.g., Oke 1990) standard stars were imaged along with the galaxy fields on nine photometric nights for H$\\alpha$ and broad-band calibration. Most of the galaxies were calibrated on more than one night and yielded a typical photometric error of 0.06 mag for all broad bands. Typical exposure times were 20 minutes for the U and H$\\alpha$ filters, and 10 minutes for the other broad-band filters. Each galaxy was imaged at least three times through each filter, but for most galaxies 4-5 images per filter were obtained. Thus, at least 30 minutes of integrated exposure time were collected through the B, V, R and I filters, and at least 60 minutes through the U and H$\\alpha$ filters. The WiseObs set of filters was replaced between the two sessions - fall of 2006 and fall of 2007. Due to instrumental problems we discarded all U-band images taken in 2006. Three galaxies were imaged and calibrated through the new U-band filter in 2007/8. Note also that all long-exposure images in the I band showed interference fringes. These were removed by an interactive script that uses the IRAF \\textit{rmfringe} command and an image of the fringe pattern with no objects in it, created by median-combining many night-time I-band images of different sky fields. The H$\\alpha$ filter used has $\\lambda_{c}=6559\\pm$2{\\AA}, FWHM=56$\\pm$2{\\AA}, peak transmittance 58.9$\\pm0.3\\%$, and out-of-band transmittance $\\leq$0.06\\% (Spector 2006). Note that the measured H$\\alpha$ line flux includes, in principle, the [NII] lines at 6548{\\AA} and 6584{\\AA}. Since the ratio of [NII] to H$\\alpha$ line fluxes in dwarf irregular galaxies is $\\leq$10\\% (e.g, Kennicutt 1983), we disregard the possible [NII] contribution through the $H\\alpha$ filter. This is specifically justified for the four objects in our sample with global spectrophotometry from Moustakas \\& Kennicutt (2006), these all have [NII]/H$\\alpha \\leq$0.08. The galaxies of the sample have radial velocities $\\leq$610 km sec$^{-1}$; this translates into a minor redward shift of the H$\\alpha$ emission line ($\\leq12${\\AA}), for an effective filter transmission of at least $\\sim95\\%$ of the peak transmittance. Since the galaxies are at $\\sim$zero redshift, the R filter was used for continuum measurement. Assuming a pessimistic photometric error of 0.1 mag, and adding in quadrature the possible [NII] contribution and the filter transmission contribution, the resulting H$\\alpha$ flux measurement accuracy is 0.15 mag. This increases to 0.18 mag for the net-H$\\alpha$ calculation (see below) due to the photometric error in the R band. For some galaxies, where both R and H$\\alpha$ have low photometric errors, the real calculated photometric error is smaller than this ``typical'' value. We collected redshifts and HI line fluxes, which were converted to mass using $M_{HI} = 2.356 \\times 10^{5}D^{2}F_{c}$ with $M_{HI}$ in solar units and $D$ in Mpc, from the ALFALFA survey (see Giovanelli et al.\\ 2005, Saintonge et al.\\ 2008) and other published data (e.g, Huchtmeier \\& Richter 1989, Karachentsev et al.\\ 2004, NED). NUV images for most of the sample galaxies were retrieved from the GALEX archive (\\textit{http://galex.stsci.edu/GR4/}). The galaxies were measured with the same aperture as the broad-band images, and the measurements were calibrated using the formal GALEX procedure (Morrissey et al.\\ 2007). The GALEX NUV band has an effective wavelength of 2315.7{\\AA}, a peak at 2200{\\AA}, and effective bandwidth is from 1771{\\AA} to 2831{\\AA} (Morrissey et al.\\ 2007). We assume similarity between this filter and the 2200{\\AA} filter calculated in the Bruzual \\& Charlot models (2003b) and use the derived NUV magnitudes to construct a (NUV-V) colour index. The typical photometric error of this colour index is similar to those of the other indices used here. \\subsection {Galaxy photometry} The galaxies were measured using the IRAF ``Ellipse'' function at a surface brightness level of $\\mu_B \\sim25.5$ mag arcsec$^{-2}$. The sky background level was determined using a galaxy-dependent method. For each image, the sky background was measured in an elliptical annulus whose inner semi-major axis (SMA) was $\\sim$1.35 times the primary SMA and its outer SMA was $\\sim$1.75 times the primary SMA. The ``Primary SMA'' term refers to the SMA of the elliptical aperture within which the galaxy was measured (at a brightness level of $\\mu_B \\sim25.5$ mag arcsec$^{-2}$). The final result for each band was corrected for extinction and for a possible colour term. The results were also corrected for Galactic extinction (GE) according to the prescriptions of Schlegel et al.\\ (1998). Since the galaxies are in a similar sky area, the typical Milky Way (MW) extinction values for this area are $A_{U}\\sim0.4$, $A_{B}\\sim0.34$, $A_{V}\\sim0.26$, $A_{R}\\sim A_{H_{\\alpha}}\\sim0.2$, and $A_{I}\\sim0.17$, but exact values were calculated specifically for each galaxy. For the GALEX NUV measurement the typical MW extinction was $\\sim0.9$ mag. The error in the determination of the extinction values is $\\sim10\\%$. The continuum flux contribution through the R filter was subtracted from the H$\\alpha$ measurement to provide the net-H$\\alpha$ (nH$\\alpha$) line flux. The derived nH$\\alpha$ was subtracted from the average $R$ flux, thus the final $R$-band magnitudes listed below are free of H$\\alpha$ line emission. The nH$\\alpha$ images were produced by subtracting the R image from the H$\\alpha$ image, after normalization by the averaged total flux of the reference stars. These nH$\\alpha$ images were used to find, count and measure individual HII regions. The photometry results, H$\\alpha$ fluxes, derived SFR (see below) and HI mass data are presented in Table~\\ref{T3}. We compared our derived photometry with published values and found reasonable compatibility. Four of our objects were observed by Moustakas \\& Kennicutt (2006) in their scanning-slit spectrophotometric survey of spiral galaxies. For these objects we found also reasonable correspondence with the Moustakas \\& Kennicutt values for the H$\\alpha$ flux being, on average, 8\\% brighter than ours. Note that the brightest objects of the sample, NGC 672 and NGC 784, have absolute magnitudes -16.9 and -16.4 if located at a distance of 5 Mpc, and could be considered ``dwarf galaxies'' even if at twice the distance. This demonstrates that the sample is indeed composed exclusively of low-luminosity galaxies. \\subsection {Comparison to galaxy evolution models} The SFH of each galaxy was derived by comparing the observed global colours to predictions of galaxy evolution models (GEMs). The GEMs are based on the GALAXEV library (Bruzual \\& Charlot 2003a, 2003b). Since the main SF characteristic in dwarf galaxies is that it is generally proceeding in bursts, we selected to compare our observational results to model predictions of stellar populations formed in bursts. We used the 26 default models pre-calculated in GALAXEV using the Padova 1994 evolutionary tracks: 13 models use the Salpeter (1955) IMF and 13 the Chabrier (2003) IMF. No significant differences were found between the results obtained with either of the two IMFs. The comparisons were done for the optical colours [(U-B), (B-V), (V-R) and (V-I)], as well as for the (H$\\alpha$-V) colour, which is produced in a procedure described below, and for the predicted (2200{\\AA}-V) colour assumed to be identical to the (GALEX NUV-V) colour. % Using the H$\\alpha$-V colour enhances the fit reliability since the line emission is determined by the amount of young massive stars in a galaxy. The models predict the number of LyC photons and we used this and the V magnitude to derive a distance-independent colour for each galaxy. The H$\\alpha$ luminosity is related to the LyC photon flux (Osterbrock 1989): \\begin{equation} N_{c} = 7.43 \\times 10^{11}L(H\\alpha), \\end{equation} where $L(H\\alpha)$ is in erg/s, which leads to: \\begin{equation} H\\alpha-V = 129.8 -2.5 \\times log[N_{c}] - V_{abs}, \\end{equation} where both $log[N_{c}]$ and $V_{abs}$ [mag] are calculated in the model. All models use standard lower and upper mass cutoffs of $m_{L}=0.1M_{\\odot}$ and $m_{U}=100M_{\\odot}$, and differ in metallicity, which ranges from 0.0001 to 0.05. The models calculate spectra and galactic colours at 220 time steps, from t=0 to t=20 Gyr. They were used to create a colour data base for each time step and for each metallicity. Since the models describe a single generation of stars formed in an instantaneous SF burst that only ages after being formed, a script was written to find the best $\\chi^{2}$ fit of any combination of colours and burst times from the data base. The minimal $\\chi^{2}$ criterion calculated by the script is given by: \\begin{equation} {\\rm Min}(\\chi^{2}) = {\\rm min} [\\sum (\\Delta colour)^{2}/err^{2}] / (N_{colours} - N_{bursts}). \\end{equation} where $\\Delta colour$ is the difference between each measured colour and the linearly-combined model-produced colour, $err$ is the error of the measured colour, and the sum refers to the measured colours (NUV-V, U-B, B-V, V-R, V-I and H$\\alpha$-V). $N_{colours}$ is the number of available colours and $N_{bursts}$ is the number of bursts in the allowed combination. Here we used primarily a combination of two bursts and the variables are the percentage of light produced by each population, the metallicity of this population, and the time of the burst. Since we fit essentially five predicted variables (fraction of light from the first burst in the observed brightness, the times of each of the two bursts, and the metallicity of each burst) to five measured colours [(U-B), (B-V), (V-R), (R-I) and (H$\\alpha$-V)], the fit should be, in principle, fully determined. For cases where we have also (NUV-V), the fit to two bursts should be overdetermined. Instances where we lack U-band photometry cause the fit to be under-determined. For some cases we were forced to assume more than two SF bursts; it is probable that these fit results are not unique. The script outputs not only the best fit, but also the next nine fits to determine the spread of the multiple solutions. The dispersion of the nine additional results around the best fit, in terms of the burst times, burst weight, and metallicities, was used as the error estimate. The dispersion around the best fit was typically 1-3 Myr for the very recent bursts, whereas for the earlier bursts that occurred a few Gyr ago it was $\\lesssim$1 Gyr. For very ``old'' bursts, $\\sim$10 Gyr, the dispersion is high and the time resolution is low; we therefore indicate that these SF bursts took place at least 10 Gyr ago without specifying exactly when. In addition, there are sometimes two different possibilities described by the 10 best fits, which have similar $\\chi^{2}$ values and we list both options below. The dispersion of the relative weights of the bursts (fraction of light measured now) is $\\sim10\\%$ around the best fit values. The least constrained variable in the fitting procedure is the metallicity, since the observed colours can be constructed from bursts at two specific times, but with different metallicities. For a wide range of observed colours, different metallicity models shows differences smaller than the observational error. For this reason, the best fit sometimes indicates a galaxy metallicity decreasing with time and such apparently non-physical solutions are retained, since they could hint at an influx of extragalactic low-metallicity gas. ", "conclusions": "The SFH of the galaxies was derived by comparing the measured galactic colours to GEMs, and the results are summarized in Table \\ref{T2}. Most sample galaxies were apparently formed more than 10 Gyr ago, with some possibly forming somewhat later (6-10 Gyr ago). All galaxies, with the exception of AGC 122834, experienced a recent SF burst at similar epochs, $\\sim$1 to a few tens of Myr ago. This could be interpreted as a result of interactions that were at least partially responsible for the SF bursts, or as the influence of an SF mechanism that might be at work synchronizing the SF in separate and non-interacting galaxies. AGC 122834 does not seem to have experienced a significant recent burst; this could imply that it is not a member of the NGC 784 group, or that it is sufficiently distant from the other galaxies not to be affected by whatever mechanism is synchronizing the SF in the other objects. \\begin{figure*}[ht] \\centering{ \\includegraphics[width=12cm]{F_SFRVsMHI.eps} \\caption{The relation between SFR and HI mass found for the objects studied here is similar to that measured in other environments for late-type galaxies.} \\label{fig:SFRHI}} \\end{figure*} \\begin{figure*}[ht] \\centering{ \\includegraphics[width=12cm]{F_SFRVsMB.eps} \\caption{Relation between SFR and absolute B magnitude. A similar relation was found for galaxies in the nearby Universe (e.g., Karachentsev \\& Kaisin 2007)} \\label{fig:SFRVsmB}} \\end{figure*} \\begin{figure*}[ht] \\centering{ \\includegraphics[width=12cm]{F_SFRVsD.eps} \\caption{SFR vs. distance from the main galaxies. In this plot we consider the distances of each galaxy from the main galaxy in its group.} \\label{fig:SFRvsDIST}} \\end{figure*} Dwarf galaxies are common in the immediate vicinity of larger galaxies (Miller 1996, Cote et al.\\ 1997). Some are believed to form as results of strong interactions or mergers between larger galaxies (e.g., Hunsberger et al.\\ 1996, Bournaud \\& Duc 2006) and can also be explained by hierarchical clustering scenarios that sometimes include supernovae winds (e.g., Dekel \\& Silk 1986, Nagashami \\& Yoshii 2004). Models can properly explain their observed colours only if the stars in these galaxies were formed in short bursts separated by long quiescence periods (e.g., Tosi et al.\\ 1991 and references therein). The recent bursts could possibly be caused by the accretion of gas retained in the dark halos of these dwarfs (Dekel \\& Silk 1986) or collected from intergalactic space (see below). % We tested correlations of the individual SFRs against various parameters. Figure~\\ref{fig:SFRHI} shows the relation between the SFR and the HI mass. The galaxies, though belonging to the same groups, show a variety of SFRs; this was found for other galaxy groups (e.g., Karachentsev \\& Kaisin 2007). We found a linear correlation of $log[SFR]$ with $log[HI_{mass}]$, which implies that the more material is available for SF, the more intense the SF would be. Kennicutt (1998), Taylor (2006) and others found that irregular and spiral galaxies follow the relation [SFR]$\\propto M_{HI}^{1.4}$ (see also Karachentsev \\& Kaisin 2007). The galaxies here, as Figure \\ref{fig:SFRHI} shows, follow a relation with a similar slope of 1.3$\\pm0.15$. % We also found that $log[SFR]$ correlates with M$_B$, both derived assuming only Hubble expansion; see Figure~\\ref{fig:SFRVsmB}). This indicates that the more intense the SF is, the more significant is the existing, relatively young, stellar population. Karachentsev \\& Kaisin (2007) found a similar result for 150 galaxies in the local Universe where the relation was [SFR]$\\propto$L$_{B}$. The galaxies in our sample follow a similar linear relation (Figure \\ref{fig:SFRVsmB}) with a slope of $\\sim$-0.4$\\pm$0.04, which corresponds to [SFR]$\\propto L_{B}^{1\\pm0.1}$, similar to that found by Karachentsev \\& Kaisin. The SFR behaviour vs. the projected distance from the main galaxies of each group is shown in Figure~\\ref{fig:SFRvsDIST}. The underlying assumption here is that the DGs might be formed in tidal interactions among the major galaxies in each group, thus the further away a DG would be from its ``main progenitor'' major galaxy, the weaker should its SF process be. It is difficult to evaluate the role of interactions in the triggering of SF in these groups, primarily since we do not have accurate 3D distances. The SFR is highest in the interacting pair NGC 672/IC 1727. The Sdm galaxies NGC 784 and UGC 1281, which might also be interacting, show relatively high SFR values as well. However, as Figure~\\ref{fig:SFRvsDIST} shows, the six other DGs do not show a steadily decreasing SFR with (projected) distance from their ``main progenitors'', but rather a $\\sim$uniform and low SF of $\\sim10^{-4}$ M$_{\\odot}$/yr. More isolated galaxies (that are also dIrr) show independently high SFRs. We interpret this behaviour, along with the similar and recent SF burst times, as possible evidence for synchronicity in star formation in galaxies formed along a dark matter (DM) filament threading a nearby void. Brosch et al.\\ (2004) found that galaxy interactions in dIrr galaxies are probably not a primary SF trigger and this seems to be the case here. The recent SF bursts in the dIrr galaxies of our sample, all taking place in the last few tens of Myrs, could be caused by accretion of extragalactic gas onto most of the galaxies. The accreted gas then forms stars upon collapse onto the disks. The lack of dependence of the SFR on the (projected) distance from the brightest and most massive galaxies, and the timing similarity of the recent bursts, show that the potential main disturbers have minimal or nil influence on the dIrr galaxies in their vicinity. Another element, such as the postulated influx of extragalactic gas, could probably be the trigger of the synchronous SF. At this point one remark is in order regarding our modeling the SF process as a collection of instantaneous bursts, separated by long quiescent intervals. The justification for this was that this SF mode is the one recognized to take place in dwarf galaxies, and all the objects here are dwarfs. However, if we examine the HI depletion time due to the star formation, by dividing the total HI mass of each object by its calculated SFR, we find that all objects could sustain SF at an almost constant level for long periods indeed, of order 10$^9$ yrs. Note that we have no information about the HI distribution; this could be very extended given the size of the ALFALFA beam, with only part of the HI taking part in the SF process. The long time needed to exhaust the HI reservoir could indicate that a quasi-continuous and low-level SF process might also explain the results. We emphasize, however, that the important issue here is that almost all objects show present-day SF. The described behaviour and the proposed interpretation are, after all, not that surprising. It has been known that only strong interactions at short separations between galaxies of comparable masses enhance SF, and that the SFR increases as the strongly interacting galaxies come closer (e.g., Icke 1985). This could perhaps be the case in the NGC 672/IC 1727 (VV338) pair, but not in most of the galaxies studied here. In addition, it has been known that fairly isolated galaxies (in relatively denser environments) show generally higher SFR the more isolated they are. This could perhaps be the case for NGC 855 and UGC 1561. % Kere\\v{s} et al. (2005) showed, using hydrodynamic simulations, that low-mass galaxies with baryonic mass M$_{gal}\\leq10^{10.3}$M$_{\\odot}$ (or halo mass M$_{halo}\\leq 10^{11.4}$M$_{\\odot}$) can accrete intergalactic gas in a ``cold mode'', with the accreted gas colder than 10$^5$K thus capable of directly forming stars. The cold accretion in the simulations is often directed along DM filaments, allowing galaxies to efficiently draw gas from large distances. Such DM filaments form in N-body simulations such as those presented by Hahn et al. (2007), where it was also shown that that haloes in filaments are more oblate than cluster halos at high masses and that haloes in filaments tend to have their spin vectors pointing along the filaments. The prediction regarding the spin alignment result of Hahn et al. (2007) is that the major axes of disky galaxies in the filament discussed here should be approximately perpendicular to the line along which the galaxies appear to be arranged. We checked this by estimating the position angles of the major axes of the galaxies (listed in Table~\\ref{gal}) and comparing them to the general direction of the projected filament. The distribution of PA values shows two preferred values: one, with seven galaxies is centered at PA$\\simeq$305$^{\\circ}\\pm12^{\\circ}$. The other peak is broader and is centered at PA$\\simeq$34$^{\\circ}\\pm31^{\\circ}$. The general direction of the galaxy filament, using the same convention of measuring PAs, is $\\sim$303$^{\\circ}$. The conclusion is therefore that the Hahn et al. (2007) prediction regarding the correlated galaxy spins is fulfilled in a mixed-up way for this nearby galaxy alignment, with seven objects having their spin approximately aligned with the general direction of the filament and the other objects with their spin axes approximately perpendicular to the filament direction. Dekel \\& Birnboim (2006) showed that low-mass galaxies are built by cold gas streams in haloes below a critical shock-heating mass M$_{shock}\\leq10^{12}$M$_{\\odot}$, not by intergalactic gas heated by a virial shock. This accretion mode yields efficient star formation in low-mass galaxies. The only ingredient then needed to understand the galaxies studied here is to assume that the phenomenon can take place not only at z$\\geq$2, but also at z$\\approx$0. We could, therefore, assume that most of the 14 galaxies studied here do line up along a DM filament in a nearby void, and that they all exhibit a present-day synchronous star formation burst triggered by cold gas accretion from intergalactic space that is being focused by the DM filament. We presented U, B, V, R, I, H$\\alpha$ and NUV photometric measurements of 14 fairly isolated galaxies in the local Universe. The galaxies are in the same sky area, most are arranged on a linear configuration, and have radial velocities $\\leq$610 km sec$^{-1}$. 11 of these galaxies, and one HI cloud with no optical counterpart, were previously assigned to two groups of galaxies. All visible galaxies are ``dwarfs'' with M$_B\\geq$--18 and can be morphologically classified as ``very late type'', including the dominating objects in each group. Three other galaxies were often considered to be field galaxies but are similar in absolute magnitude and classification to the other objects. The objects appear to form a single kinematically well-behaved ensemble that does not separate naturally into two groups. We derived the SFR and SFH of each object using photometry in the various bands and examined these properties in the context of the galaxy environment. The galaxies show relations between the SFR and the HI mass, or m$_B$, known for field galaxies. The SFHs imply that similar low-intensity SF bursts took place in most galaxies a few Myr ago, despite their being spread along a $\\geq$0.5 Mpc long feature. The behaviour of the SFRs, which do not decline steadily with increasing distance from the brightest galaxies in each group, indicates that interactions do not affect the SF in these galaxies to the extent one would expect from strongly interacting galaxies. In particular, no signs of strong gravitational interactions, such as tidal tails or global disturbances, are observed. The main galaxies do appear to be interacting but on a one-to-one basis and do indeed show enhanced SF. Other dIrr galaxies in the same vicinity show only low SFRs. More isolated galaxies show independently high SFRs. We propose that the observational evidence argues in favor of interpreting the galaxies as located on a DM filament that is itself located in a low-galaxy-density region, and is accreting intergalactic cold gas focused by the filament. We are therefore witnessing in the classical NGC 672 and NGC 784 groups of relatively bright and nearby late-type galaxies the basic phenomenon of hierarchical clustering, the direct formation and growth of small galaxies out of intergalactic gas accreted on a dark matter ``backbone''. This nearby galaxy collection offers, therefore, an ideal opportunity to study the phenomenon of hierarchical clustering in significant detail. One possible conclusion from our proposed interpretation is that the gas being accreted could be observable in the vicinity of the galaxies; a way to detect it would be through very low column density detection of Lyman $\\alpha$ absorption lines at the redshifts of the galaxies in spectra of background QSOs. Another would be to search for diffuse HI to much lower limits than achieved during the ALFALFA survey; it is unlikely that an extension of the ALFALFA scans will bring useful returns and perhaps this is a project that should wait for the entrance in operation of SKA." }, "0808/0808.3573_arXiv.txt": { "abstract": "{Near-infrared (hereafter NIR) data may provide complementary information to the traditional optical population synthesis analysis of unresolved stellar populations because the spectral energy distribution of the galaxies in the 1-2.5\\,$\\mu$m range is dominated by different types of stars than at optical wavelengths. Furthermore, NIR data are subjected to less absorption and hence could constrain the stellar populations in dust-obscured galaxies.} {We want to develop observational constraints on the stellar populations of unresolved stellar systems in the NIR.} {To achieve this goal we need a benchmark sample of NIR spectra of ``simple'' early-type galaxies, to be used for testing and calibrating the outputs of population synthesis models. We obtained low-resolution (R$\\sim$1000) long-slit spectra between 1.5 and 2.4\\,$\\mu$m for 14 nearby early-type galaxies using SofI at the ESO 3.5-m New Technology Telescope and higher resolution (R$\\sim$3000) long-slit spectra, centered at the Mg{\\sc I} at $\\sim$1.51\\,$\\mu$m for a heterogeneous sample of 5 nearby galaxies observed with ISAAC at Antu, one of the 8.2-m ESO Very Large Telescope.} {We defined spectral indices for CO, Na{\\sc I}, Ca{\\sc I} and Mg{\\sc I} features and measured the strengths of these features in the sample galaxies. We defined a new global NIR metallicity index, suitable for abundance measurements in low-resolution spectra. Finally, we present an average NIR spectrum of an early-type galaxy, built from a homogenized subset of our sample.} {The NIR spectra of the sample galaxies show great similarity and the strength of some features does correlate with the iron abundance [Fe/H] and optical metal features of the galaxies. The data suggest that the NIR metal features, in combination with a hydrogen absorption feature may be able to break the age-metallicity degeneracy just like the Mg and Fe features in the optical wavelength range.} ", "introduction": "State-of-the-art space-based instrumentation can only resolve galaxies from the Local Group well and in the more distant objects we usually can see only the tip of the red giant branch stars which is rarely sufficient for population analysis. This leaves us with the difficult task of trying to recover the stellar population of unresolved galaxies from their integrated properties. The complex mix of stellar populations found in most of them usually makes it possible to constrain only the most recent generation of stars. However, such properties as the stellar kinematics and the present-day metal content are dominated by the overall star formation history, and together with the well known age-metallicity degeneracy \\citep{fab73,oco86,wor94} they often lead to non-unique solutions for the stellar populations. In this respect, the Lick/IDS system of indices pioneered by \\citet{bur84} and \\citet{fab85}, and developed further by \\citet{tra98, tra00, tra05}, has been particularly successful for interpreting the integrated optical light of galaxies. However, new constraints are necessary to interpret more complicated systems and one possibility is to widen the spectral range towards the near-infrared (NIR) because the light in different wavebands is dominated by different populations of stars. The NIR passbands are dominated by light from older and redder star and therefore offer us the possibility to study other stellar populations than is possible with optical spectra alone. In addition, abundance determinations through optical spectroscopy are not possible for heavily reddened evolved stellar populations such as dusty spheroids or some bulge globular clusters hidden by dust \\citep[e.g.,][]{ste04}. NIR spectroscopic observations could overcome these problem because the extinction in the $K$-band is only one-tenth of that in the $V$-band. Most of the previous work at NIR was focused on either active galactic nuclei (AGNs) or objects with very strong star formation, including recent surveys of ultra-luminous infrared galaxies \\citep{gol95,mur99,mur01,bur01}, luminous infrared galaxies \\citep{gol97,reu07}, starbursts \\citep{eng97,coz01}, Seyfert galaxies \\citep{iva00,sos01,reu02,boi02}, LINERs \\citep{lar98,alo00,sos01}, and interacting galaxies \\citep{van97,van98}. Relatively few NIR spectroscopic observations exist for ``normal'' galaxies. Only \\citet{man01} provided low-resolution ($R\\sim400$) template spectra for galaxies of different Hubble types, including some giant ellipticals. Such data, together with the corresponding analysis is a necessary first step towards developing a system of spectral diagnostics in the NIR because well-understood galaxies with relatively simple star forming history will allow us to tune the NIR population synthesis models. Recently, \\citet[][hereafter S08]{sil08} studied the stellar populations of eleven early-type galaxies in the nearby Fornax cluster by means of $K$-band spectroscopy. A few prominent NIR features were first studied by \\citet{ori93} who demonstrated that they represent a superb set of indicators for constraining the average spectral type and luminosity class of cool evolved stars. This conclusion was later confirmed by \\citet{for00} and \\citet{iva04}. Furthermore, the same NIR features appear to be promising abundance indicators \\citep{fro01,ste04}. Here we describe two new data sets of high quality NIR spectra of ellipticals/spirals designed to provide a benchmark for future NIR studies of unresolved galaxies: {\\it (i)} low-resolution spectra covering the range from 1.5 to 2.4\\,$\\mu$m that include strong features such as CO, Na{\\sc I} and Ca{\\sc I} that are traditionally studied, and {\\it (ii)} moderate resolution spectra around the Mg{\\sc I} absorption feature at 1.51\\,$\\mu$m. This is the second strongest Mg feature in the $H$- and $K$-band atmospheric windows (after the feature at 1.71\\,$\\mu$m) and at zero redshift it is located in a region of poor atmospheric transmission, making it difficult to observe in stars. However, the redshift of external galaxies moves it into a more transparent region \\citep{iva01}, just the opposite of the 1.71\\,$\\mu$m Mg{\\sc I} which becomes affected by the red edge of the $H$-band atmospheric window. We are only few years from the launch of the James Webb Space Telescope \\citep{gar06}, a space-based infrared telescope that will have unprecedented capabilities. Therefore, in the near future the application of NIR spectroscopy to the study of galaxy properties will be limited not by lack of data but by our understanding of spectral features at these wavelengths. Improving the characterization of the NIR indices is a timely step in this direction. The paper is organized as follows. The sample selection is discussed in Sect.\\,\\ref{sec:Sample}. The NIR spectroscopic observations and data reduction are described in Sect.\\,\\ref{sec:Obs}. The definition of the new NIR spectral indices and their measurements are given in Sect.\\,\\ref{sec:Indices}. Results are discussed in Sect.\\,\\ref{sec:Discus} and summarized in Sect.\\,\\ref{sec:Results}. ", "conclusions": "} \\subsection{General appearance of the spectra} The spectra of our sample galaxies appear qualitatively similar in most of the NIR features except for a large spread in the Na{\\sc I} values is present. This similarity is not surprising because we selected mostly giant ellipticals and spheroids, which are all relatively metal rich, have no significant recent star formation, and only weak nuclear activity, if any. This conclusion agrees with \\citet{man01} who demonstrated that galaxies within the same Hubble type have nearly identical NIR spectra. On the other hand, the large spread of NaI, with respect to the observational errors, appears to be real, and suggests variety of star formation and enrichment histories among the galaxies in our sample. Two of the three galaxies with systematically weaker NaI exhibit strong H$\\beta$, indicating that the weak NaI might be related to the presence of younger stellar populations (the third galaxy lacks optical spectroscopy). The strongest features are the CO absorption bands in both the $H$ and $K$-band atmospheric windows. They originate in K and M stars, as can be seen from the libraries of stellar spectra described in \\citet{lan92}, \\citet{ori93}, \\citet{dal96}, \\citet{mey98}, \\citet{wal97} and \\citet{for00}. A number of weaker metal absorption features are visible as well: Si{\\sc I} at 1.589\\,$\\mu$m, Mg{\\sc I} at 1.711\\,$\\mu$m, Na{\\sc I} at 2.206 and 2.209\\,$\\mu$m and Ca{\\sc I} at 2.261, 2.263, and 2.266\\,$\\mu$m. They are also present mostly in cool stars \\citep[i.e.][]{kle86,ori93,wal97,for00,mey98}. As mentioned above, we only consider the redder features with $\\lambda\\geq$1.65$\\mu$m. Not surprisingly, our galaxies show no Fe{\\sc II} at 1.644\\,$\\mu$m or H$_2$ lines line emission, usually associated with supernova activity and supernova related shocks, respectively. There is also no Br$\\gamma$ emission. \\subsection{Behavior of the NIR spectral indices}\\label{sec:Behaviour} Evolutionary synthesis models \\citep[i.e.][]{wor94} are the usual method to interpret the behavior of spectral features in galaxies, but the simple comparison of these features with the galaxy parameters can be informative too. For example, the optical indices, Mg$_2$\\,, $<$Fe$>$ and \\Hb show tight correlation with the central velocity dispersion \\citep[e.g.][]{ter81,ber98,meh03,mor08}, suggesting that the chemical and the dynamical evolution of ellipticals are intertwined. The EW of our NIR spectral indices, the iron abundance and metallicity are plotted versus the central velocity dispersion in Fig.\\,\\ref{fig:VelDips}. The loci of the S08 data (see their Fig. 13) are shown with dashed lines. Among the galaxies, NGC\\,5077 (\\#12) and NGC\\,6909 (\\#14) possess very weak Na{\\sc I}, Ca{\\sc I} lines, and their relative errors are significant. The correlations between the NIR indices and velocity dispersion for the sample galaxies are plotted in Fig\\,\\ref{fig:VelDips}; the fit do not consider NGC\\,4281 (\\#4), NGC\\,5077 (\\#12) and NGC\\,6909 (\\#14) because of the reasons explained further on. The correlation with CO shows a similar slope as the one found by S08 but with a different zero-point. The different definition of the index and especially the position of the continuum bandpass can lead to a systematic variation of the EW values. This effect seems to be more evident in the CO band, whose equivalent width is measured with only extrapolated continuum blue-wards of the feature. The agreement with S08 is closer in the Na{\\sc I} versus $\\sigma$ relation although our data show bigger scatter. Without pretending to trace a general conclusion we found relatively large NaI values compared to stars and something similar was also found by S08. The iron abundance [Fe/H] and the total abundance [Z/H] only show very weak correlations. Active galaxies are clustered at the high end of the velocity dispersion distribution which is understandable because their velocity dispersion measurement may be affected by the black hole and the active nucleus. Note that our highest velocity dispersion galaxy is a Type 2 Seyfert and the CO band may suffer some dilution, as discussed in \\citet{iva00}. \\begin{figure} \\includegraphics[width=9truecm]{vd02a.eps} \\caption{The equivalent widths of the NIR spectral indices, iron abundance [Fe/H], and metal abundance [$Z$/H] in the sample galaxies as a function of the central velocity dispersion. The labels correspond to the row numbers in the Table \\ref{tab:GalProperties1}. The open and filled circles refer to quiescent and active galaxies, respectively. The small and large circles refer to galaxies with an age between 5 and 10\\,Gyr and $>$10\\,Gyr, respectively. Galaxies with no known ages were assumed to be old. The dashed lines corresponds to the relations find by S08. The solid line in each panel represents the linear regression (y = ax + b) through all the data points except for \\#4, \\#12 and \\#14. The Pearson correlation coefficient (r) and the results of the linear fit are given. Typical error bars are shown.} \\label{fig:VelDips} \\end{figure} The EW of our NIR spectral indices are plotted against each other in Fig.\\,\\ref{fig:IR_IR} and compared to the relations for cluster stars and solar neighborhood giants by S08 and \\citet{dav08}, respectively. Three galaxies -- NGC\\,4281 (\\#4), NGC\\,5077 (\\#12) and NGC\\,6909 (\\#14) -- are above the relation of S08 for the cluster stars, which hints at different chemical enrichment history with respect to the rest of the sample. Unfortunately the literature lacks much data about these objects and although NGC\\,6909 has the lowest velocity dispersion in our sample, these galaxies do not stand out in any respect, including in the optical Mg$_2$ versus $\\sigma$ diagram. Further investigation of these galaxies is necessary. The rest of our objects populates a locus that follows similar trend as the galaxies of S08. In the Ca{\\sc I} versus CO plot we see a correlation with significant scatter, and with a different slope than in S08. Finally, the Mg{\\sc I} versus CO plot is dominated by scatter. Galaxies with traces of nuclear activity, evident in other wavelength ranges than the NIR, do not seem to separate from the rest of the sample, and they show no traces of emission lines. Therefore, the contribution of their AGNs is negligible with respect to the rest of the galaxy. \\begin{figure} \\includegraphics[width=9truecm]{IR_IR02a.eps} \\caption{The equivalent widths of the NIR indices plotted against each other. The symbols are the same as in Fig.\\,\\ref{fig:VelDips}. The dashed lines correspond to the relations found by S08. The dotted line corresponds to the relation by \\citet{dav08} for giant stars in the solar neighborhoods. All these relations are plotted in their observed range. The solid line represents the linear regression (y = ax + b) through all the data points except for \\#4, \\#12 and \\#14. The Pearson correlation coefficient (r) and the results of the linear fit are given. Typical error bars are shown.} \\label{fig:IR_IR} \\end{figure} Figure\\,\\ref{fig:IR_FeH} shows the EW of the NIR spectral indices versus the Mg$_2$ measurements from \\citet{ben93}. Na{\\sc I}, and to a lesser extent Ca{\\sc I} and CO, do show trends with respect to Mg$_2$, with NGC\\,4281 (\\#4), NGC\\,5077 (\\#12) and NGC\\,6909 (\\#14) standing out. The correlation of Na{\\sc I} with both $\\sigma$ and Mg$_2$ suggests that the Na{\\sc I} feature is dominated by the stellar photosphere rather than by the interstellar medium. \\begin{figure} \\includegraphics[width=9truecm]{IR_FeH02a.eps} \\caption{The equivalent widths of the NIR indices plotted against the Mg$_2$ indices measured by \\citet{ben93}. The symbols are the same as in Fig.\\,\\ref{fig:VelDips}. The solid line represents the linear regression (y = ax + b) through all the data points except for \\#4, \\#12 and \\#14. The Pearson correlation coefficient (r) and the results of the linear fit are given. Typical error bars are shown.} \\label{fig:IR_FeH} \\end{figure} Given that [Fe/H] measurements are not available for the entire sample and that Mg$_2$ is not representative of the total chemical abundance of a galaxy because of the varying abundance ratios, we attempted to create a combined Iron-and-$\\alpha$-element index similar to that defined by \\citet{gon93} and more recently by \\citet{tom03}: \\begin{equation} {\\rm [MgFe]'} = \\sqrt{{\\rm Mg}b\\times (0.72 \\times {\\rm Fe}5270+0.28 \\times {\\rm Fe}5335)} \\end{equation} Such a combined indicator is expected to decrease the effect from the varying $\\alpha$/Fe ratio. Since not all the components of this indicator are available we directly substitute the iron and magnesium indices by defining a new indicator: \\begin{equation}\\label{Eqn:FeMg2} {\\rm [MgFe]''} = \\sqrt{{\\rm Mg}_2\\times {\\rm Fe}5335} \\end{equation} The results are shown in Fig.\\,\\ref{fig:IR_MgFe}a-c. The NaI index of NGC\\,5077 (\\#12) and the CO index of NGC\\,6909 (\\#14) deviate from the main loci of the other galaxies. To address this issue we plotted the optical indices of these galaxies versus the combined [MgFe]$'$ index (Fig.\\,\\ref{fig:IR_MgFe}d). This is an analog of the typical plot \\citep[i.e.][]{wor94} that allows to disentangle the age-metallicity degeneracy: the inverse \\Hb index is roughly proportional to age while the [MgFe]$'$ is dominated by metal abundance. The two galaxies exhibit strong \\Hb which suggests that they are dominated by populations of 3\\,Gyr or younger (see for example Fig. 1 in S08). NGC\\,5077 and NGC\\,6909 are also well separated from the rest of the galaxies on the \\Hb versus Na{\\sc I} plot (Fig.\\,\\ref{fig:IR_MgFe}e) which leads us to the conclusion that the NIR indices can be used to create a similar diagnostic plot, but measuring the Br$\\gamma$ feature requires better quality data than the ones described here.\\\\ We investigated the behavior of the NIR indices in relation to the H and K-band magnitude and the H-K color but no clear correlations were found. \\begin{figure} \\includegraphics[width=9truecm]{IR_MgFe02a.eps} \\caption{(a-d) The equivalent widths of the NaI, CaI, CO and H$\\beta$ indices plotted against the [MgFe]$''$ index. (e) The equivalent widths of the H$\\beta$ index as a function of the NaI index. The symbols are the same as in Fig.\\,\\ref{fig:VelDips}. The dashed lines correspond to the relations found by S08. The solid line represents the linear regression (y = ax + b) through all the data points except for \\#4, \\#12 and \\#14. The Pearson correlation coefficient (r) and the results of the linear fit are given. Typical error bars are shown.} \\label{fig:IR_MgFe} \\end{figure} \\subsection{Combined NIR metal index} Ground based NIR spectroscopy is much more time consuming than the corresponding optical observations -- higher and variable background and detectors with worse cosmetics often require one to sacrifice either resolution or signal-to-noise to obtain the data in a reasonable amount of time. To alleviate this problem, we defined a combined spectral index of all the major $K$-band metal features: \\begin{equation} \\langle {\\rm CONaCa} \\rangle = ( {\\rm CO} + \\ion{Na}{i} + \\ion{Ca}{i} ) / 3.0 \\end{equation} Various weighting schemes were tried to minimize the scatter of the basic relations (Fig.\\,\\ref{fig:CONaCa}). However, the simple average yielded the tightest relations. The carbon and oxygen are $\\alpha$ elements, while the sodium and calcium originate in both high and low mass stars. The average value of the CO index for the sample galaxies is $\\sim$15\\,\\AA, the Na{\\sc I} is $\\sim$4.3\\,\\AA\\, and the Ca{\\sc I} is $\\sim$2.7\\,\\AA\\, which means that at least 2/3 of the new index is dominated by metals produced mostly in high mass stars, i.e. early in the history of the elliptical galaxies. The other implication is that a relatively limited amount of recent star formation could affect the new index more than it would an iron peak dominated index. The lack of iron peak features in the NIR is well known as it was pointed out in \\citet{iva01}. \\begin{figure} \\includegraphics[width=9truecm]{nir02a.eps} \\caption{The equivalent widths of the newly defined $<$CONaCa$>$ as a function of the central velocity dispersion (a), Mg$_2$ (b), H$\\beta$ (c), and [FeMg]$''$ (d) indices. The symbols are the same as in Fig.\\,\\ref{fig:VelDips}. The solid line represents the linear regression (y = ax + b) through all the data points except for \\#4, \\#12 and \\#14. The Pearson correlation coefficient (r) and the results of the linear fit are given. Typical error bars are shown.} \\label{fig:CONaCa} \\end{figure} The new index improves the correlations. For example, the Pearson correlation coefficient is $\\sim$10\\% higher for the $\\langle {\\rm CONaCa} \\rangle$ vs. Mg2 and even $\\sim$20\\% higher for the $\\langle {\\rm CONaCa} \\rangle$ vs. [MgFe]'', with respect to the respective correlations where only one IR index is used. We derived these relations using only the galaxies with known parameters, i.e. excluding the poorly studied NGC\\,4281(\\#4) and, NGC\\,5077(\\#12) and NGC\\,6909(\\#14) which systematically show peculiarities with respect to the bulk of our sample. \\subsection{Average NIR spectrum of the sample galaxies} Studies of composite stellar systems (i.e. galaxies hosting an AGN) often need to subtract the contribution of the underlying galaxy. This prompted us to create an average spectrum of the galaxies in our sample. We used a homogenized subset of eight galaxies, excluding the objects with young populations. We also excluded those with low signal-to-noise ratio. The average age of the galaxies used to create the composite spectrum is 9$\\pm$2\\,Gyr (the median is 9.5\\,Gyr) and the average [Fe/H] is 0.17$\\pm$0.13 (the median is 0.16). The composite spectrum is shown in Fig.\\,\\ref{fig:AverSpec} and the values of the spectrum are listed in Table\\,\\ref{tab:AverSpec}, together with the r.m.s. values per Angstrom. We included the $H$-band section because the artifacts caused by the spectral type mismatch of the standards are minimized by the averaging of galaxies observed at different redshifts. \\begin{figure} \\includegraphics[width=9truecm]{comb01.eps} \\caption{Average $H$- (a) and $K-$band (b) spectrum of sample galaxies normalized to unity at 22000\\ \\AA\\, (thick line). The thin lines correspond to the $\\pm1\\sigma$ confidence levels. Some of the more prominent spectral features are marked.} \\label{fig:AverSpec} \\end{figure} \\begin{table} \\caption{Average spectrum of the sample galaxies observed at NTT in arbitrary units and normalized to unity at 22000\\,\\AA. The regions with zero r.m.s. at the beginning and the at the and are covered only by one or two spectra. The full table is given only in the electronic version of the journal.} \\label{tab:AverSpec} \\begin{center} \\begin{tabular}{ccc} \\hline \\hline $\\lambda$ (\\AA) & F$_\\lambda$ & r.m.s. \\\\ \\hline 14944 & 2.217 & 0.000 \\\\ 14945 & 2.384 & 0.000 \\\\ 14946 & 2.384 & 0.000 \\\\ 14947 & 2.386 & 0.000 \\\\ 14948 & 2.130 & 0.366 \\\\ 14949 & 2.572 & 0.254 \\\\ 14950 & 2.575 & 0.251 \\\\ 14951 & 2.578 & 0.248 \\\\ 14952 & 2.583 & 0.245 \\\\ 14953 & 2.588 & 0.243 \\\\ 14954 & 2.595 & 0.241 \\\\ 14955 & 2.602 & 0.240 \\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\subsection{The Mg{\\sc I} feature at 1.51\\,$\\mu$m} \\citet{iva04} pointed out the possibility of using the Mg{\\sc I} feature at 1.51\\,$\\mu$m as an $\\alpha$-element abundance indicator in the NIR. With exploratory purposes we obtained spectra of a small and diverse set of galaxies (Table\\,\\ref{tab:log1} and Fig.\\,\\ref{fig:MgI_H_band}). To carry out quantitative analysis we defined an index (Table\\,\\ref{tab:bandDef}) and for the five galaxies we measured equivalent widths of 3.3, 5.0, 3.8, 3.9 and 4.5\\AA, respectively for Mrk\\,1055, NGC\\,1144, NGC\\,1362, NGC\\,4472 and NGC\\,7714. The typical uncertainty is $\\sim0.3$\\AA. No corrections for velocity dispersion were applied. Given the diverse nature of the objects, we refrain from drawing any conclusions but we note that the relation between the optical and the NIR Mg features is not straightforward because the only two Mg$_2$ values that we have from the literature are inversely proportional to our measurements for NGC\\,1362 and NGC\\,4472. Nevertheless, these observations prove that it is feasible to measure the NIR Mg{\\sc I} feature at 1.51\\,$\\mu$m and give a basis for future NIR synthetic spectral modeling. \\begin{figure} \\includegraphics[width=9truecm]{spe04.eps} \\caption{Spectra of the galaxies observed with ISAAC in the region of the Mg{\\sc I} (1.51 $\\mu$m). The central wavelengths of the individual Mg{\\sc I} lines are shown with vertical solid lines. The passbands for the feature and the adjacent continua are shaded. The spectra were normalized to unity and shifted vertically by 0.5 for display purposes. } \\label{fig:MgI_H_band} \\end{figure}" }, "0808/0808.2480_arXiv.txt": { "abstract": "We have determined Al, $\\alpha$, Fe--peak, and neutron capture elemental abundances for five red giant branch (RGB) stars in the Galactic globular cluster M10. Abundances were determined using equivalent width analyses of moderate resolution (R$\\sim$15,000) spectra obtained with the Hydra multifiber positioner and bench spectrograph on the WIYN telescope. The data sample the upper RGB from the luminosity level near the horizontal branch to about 0.5 mag below the RGB tip. We find in agreement with previous studies that M10 is moderately metal--poor with [Fe/H]=--1.45 ($\\sigma$=0.04). All stars appear enhanced in Al with $\\langle$[Al/Fe]$\\rangle$=+0.33 ($\\sigma$=0.19), but no stars have [Al/Fe]$\\ga$+0.55. We find the $\\alpha$ elements to be enhanced by +0.20 to +0.40 dex and the Fe--peak elements to have [el/Fe]$\\sim$0, which are consistent with predictions from type II SNe ejecta. Additionally, the cluster appears to be r--process rich with $\\langle$[Eu/La]$\\rangle$=+0.41. ", "introduction": "Although few chemical analysis studies of M10 exist, the general consensus is that this cluster exhibits all of the classical characteristics observed in other Galactic globular clusters. With a metallicity of [Fe/H]$\\approx$--1.5 (Kraft et al. 1995), M10 lies near the median metallicity distribution for halo globular clusters (Laird et al. 1988). Small sample (N$\\la$15) analyses of red giant branch (RGB) stars in this cluster have revealed it to have [$\\alpha$/Fe]$\\sim$+0.30 and [el/Fe]$\\sim$0 for Fe--peak elements (Kraft et al. 1995; Mishenina et al. 2003). These values are consistent with the current generation of M10 stars having been polluted by the ejecta of type II supernovae (SNe) without significant contributions from type Ia SNe. While it has long been known that nearly all globular cluster giants show star--to--star variations of the light elements (A$\\la$27), the source of many of these anomalies has yet to be determined. Numerous observations of globular cluster stars from the main sequence to above the RGB luminosity bump have revealed declining [C/Fe] and increasing [N/Fe] ratios as a function of increasing luminosity (e.g., see reviews by Kraft et al. 1994; Gratton et al. 2004; Carretta 2008). These observations show clear evidence of CN--cycle products being brought to the surface and are a confirmation of first dredge--up predictions (Iben 1964). Smith \\& Fulbright (1997) and Smith et al. (2005) have verified this trend in M10 as well as a CN band anticorrelation with [O/Fe] for stars at various RGB luminosities. However, the large spread in [N/Fe] of about 1.0 dex found by Smith et al. (2005) in M10 stars may be evidence for primordial variations superimposed on in situ mixing. The C and N abundance anomalies are known to exist in both globular cluster and field giants, but that likeness does not extend to the well documented O/Na, Mg/Al, and O/F anticorrelations and Na/Al correlation seen solely in globular cluster stars (e.g., Gratton et al. 2004). These abundance relationships are clear signs of proton--capture nucleosynthesis, but where these processes are operating is still a mystery. Kraft et al. (1995) examined the O/Na anticorrelations of 15 bright giants in M10 along with M3 and M13, which are all globular clusters of similar metallicity ([Fe/H]$\\approx$--1.5), because M10 and M13 have extremely blue horizontal branches (HB) but M3 has a uniform distribution of blue HB, RR Lyrae, and red HB stars. The study showed that M10 appears to be an intermediate case in terms of O depletion and Na enhancement in that the average [O/Fe] is lower in M10 than in M3, but no M10 giants were super O--poor (i.e., [O/Fe]$<$--0.6), suggesting the process driving O depletion does not itself determine HB morphology. In this paper we have examined five additional RGB stars in M10 that are located above the luminosity of the horizontal branch but below the RGB tip. We have derived Al, $\\alpha$, Fe--peak, and heavy element abundances to examine how M10 fits into context with other globular clusters of similar metallicity and HB morphology. ", "conclusions": "\\subsection{Al Abundances} We have determined at least upper limits of [Al/Fe] for five giants with the cluster having $\\langle$[Al/Fe]$\\rangle$=+0.33 ($\\sigma$=0.19) and a full range of 0.50 dex. Both the star--to--star dispersion and average [Al/Fe] ratios are in agreement with observations of other Galactic globular clusters of similar metallicity (e.g., Kraft et al. 1998; Sneden et al. 2004; Cohen \\& Mel{\\'e}ndez 2005; Johnson et al. 2005; Yong et al. 2005); however, the highest [Al/Fe] ratio found in our sample is about a factor of three smaller than the $>$+1.0 dex ratios observed in M3 and M13 (Pilachowski et al. 1996; Sneden et al. 2004; Johnson et al. 2005; Cohen \\& Mel{\\'e}ndez 2005), which possess similar metallicity and, in the case of M13, a similar HB morphology. This may be due to our small sample size coupled with observations of stars well below the RGB tip, where additional Al enhancement due to extra in situ mixing may be operating (e.g., Denissenkov \\& VandenBerg 2003). Kraft et al. (1995) found M10 to be an intermediate case between M3 and M13 with regard to the amount of O depletion and Na enhancement and therefore given the likely Na--Al correlation present in this cluster one would not expect [Al/Fe] values much greater than about +0.80 dex. A complete list of our determined abundances for Al and all other elements is provided in Table 3. It has been shown that [Fe/H] determinations based on Fe I lines in metal--poor stars suffer from larger LTE departure effects than their metal--rich counterparts because of overionization due to decreased UV line blocking (e.g., see review by Asplund 2005). Correcting for this effect would drive the [Fe/H] abundance up, perhaps by as much as $\\sim$+0.30 dex at [Fe/H]=--3 (Th{\\'e}venin \\& Idiart 1999, but see also Gratton et al. 1999; Kraft \\& Ivans 2003), and thus decrease the derived [Al/Fe] ratio found here. While a few NLTE studies for Al exist (e.g., Gehren et al. 2004; Andrievsky et al. 2008) finding offsets of order a few tenths of a dex, the actual Al NLTE correction for stars in the metallicity and luminosity regime studied here are mostly unknown. Fortunately, our sample does not vary widely in either metallicity or luminosity and any NLTE corrections are likely to be very similar, suggesting at least the relative star--to--star dispersion is a real effect. In Figure \\ref{f2} we compare abundances of various elements in M10 versus those in the similar cluster M12. The [Al/Fe] abundances for both clusters are comparable and each displays a modest star--to--star dispersion. Given that the scatter is about a factor of two larger than those observed in the Fe--peak and $\\alpha$ elements, it is likely that the Al distribution is real and not an artifact of observational uncertainty. To see how M10 fits into the context of other Galactic globular clusters, we have plotted [Al/Fe] as a function of both horizontal branch ratio (HBR) and galactocentric distance (R$_{\\rm GC}$) for M10 and seven other clusters in Figure \\ref{f3}. The top panel suggests there is no significant relation between HBR and either the average [Al/Fe] ratio or the star--to--star dispersion. However, it should be noted that Carretta et al. (2007) do find a relationship between the extent of O/Mg depletions and Na/Al enhancements and the maximum temperature of stars located on the zero--age HB. The bottom panel may indicate a trend of increasing cluster average [Al/Fe] with increasing galactocentric distance; however, the sample size for each cluster varies between less than 10 to nearly 100 stars. Consequently, M10 does not appear to exhibit anomalous [Al/Fe] ratios compared to other globular clusters. \\subsection{$\\alpha$, Fe--Peak, and Heavy Elements} Nearly all globular clusters with [Fe/H]$<$--1 have [$\\alpha$/Fe]$\\sim$+0.30 to +0.50, solar Fe--peak to Fe ratios, and are r--process rich (e.g., Gratton et al. 2004). The star--to--star scatter present is usually $\\la$0.10 dex for the $\\alpha$ and Fe--peak elements and $\\sim$0.30--0.50 dex for the neutron capture elements, which is still significantly less than the 0.50--1.00 dex variations seen in light elements such as O, Na, and Al. In M10 we find the expected enhancement and small star--to--star dispersion of the two $\\alpha$ elements Ca and Ti with $\\langle$[Ca/Fe]$\\rangle$=+0.42 ($\\sigma$=0.12) and $\\langle$[Ti/Fe]$\\rangle$=+0.24 ($\\sigma$=0.06). These values are consistent with the results from Kraft et al. (1995) that found $\\langle$[Ca/Fe]$\\rangle$=+0.29 ($\\sigma$=0.07) and $\\langle$[Ti/Fe]$\\rangle$=+0.21 ($\\sigma$=0.12) for a set of 10 other upper RGB stars in this cluster. The proxy Fe--peak elements Sc and Ni exhibit near solar abundance ratios in all stars with cluster average values of $\\langle$[Sc/Fe]$\\rangle$=+0.03 ($\\sigma$=0.19) and $\\langle$[Ni/Fe]$\\rangle$=+0.09 ($\\sigma$=0.06), which are roughly consistent with Kraft et al. (1995). These abundances patterns are mirrored in M12 (see Figure \\ref{f2}), but with M10 showing a smaller range of [Cr/Fe] and [Co/Fe] abundances. The combination of $\\alpha$ enhancement and near solar Fe--peak ratios is consistent with this cluster being primarily polluted by the ejecta of type II SNe (e.g., Woosley \\& Weaver 1995). For stars near M10's metallicity, La is produced primarily via the s--process in $\\sim$1--3 M$_{\\odot}$ stars and Eu from the r--process in $\\sim$8--10 M$_{\\odot}$ stars (e.g., Busso et al. 1999; Truran et al. 2002). Our derived La and Eu abundances are consistent with the picture of massive stars producing most of the heavy elements in this cluster with $\\langle$[La/Fe]$\\rangle$=+0.08 ($\\sigma$=0.29) and $\\langle$[Eu/Fe]$\\rangle$=+0.54 ($\\sigma$=0.10). Comparing the ratio of r-- to s--process elements gives [Eu/La]=+0.41 and implies M10 is slightly more r--process rich than the average globular cluster. However, this value is within the 1$\\sigma$ range of $\\langle$[Eu/Ba,La]$\\rangle$=+0.23 ($\\sigma$=0.21) found by Gratton et al. (2004) after combining data from the literature on 28 globular clusters. A larger sample size of M10 stars is likely to decrease the star--to--star scatter observed in our La and Eu sample but will probably not change the result that the cluster is r--process rich." }, "0808/0808.3353_arXiv.txt": { "abstract": "{The peculiar hot star \\tc\\ in the open cluster IC\\,2602 is a blue straggler as well as a single-line binary of short period (2.2d).} {Its high-energy properties are not well known, though X-rays can provide useful constraints on the energetic processes at work in binaries as well as in peculiar, single objects. } {We present the analysis of a 50\\,ks exposure taken with the \\xmm\\ observatory. It provides medium as well as high-resolution spectroscopy. } {Our high-resolution spectroscopy analysis reveals a very soft spectrum with multiple temperature components (1--6\\,MK) and an X-ray flux slightly below the `canonical' value ($\\log[L_X(0.1-10.)/L_{BOL}]\\sim-7$). The X-ray lines appear surprisingly narrow and unshifted, reminiscent of those of $\\beta$\\,Cru and $\\tau$\\,Sco. Their relative intensities confirm the anomalous abundances detected in the optical domain (C strongly depleted, N strongly enriched, O slightly depleted). In addition, the X-ray data favor a slight depletion in neon and iron, but they are less conclusive for the magnesium abundance (solar-like?). While no significant changes occur during the \\xmm\\ observation, variability in the X-ray domain is detected on the long-term range. The formation radius of the X-ray emission is loosely constrained to $<$5\\,R$_{\\odot}$, which allows for a range of models (wind-shock, corona, magnetic confinement,...) though not all of them can be reconciled with the softness of the spectrum and the narrowness of the lines.} {} ", "introduction": "\\tc\\ (=HD\\,93030) is a luminous star of type B0.2V belonging to the open cluster IC\\,2602, situated at 152\\,pc \\citep{rob99}. The cluster is 30\\,Myr old, and the massive, hot \\tc\\ therefore appears to be a rare example of blue straggler. Its optical spectrum display hints of enhanced nitrogen (a three-fold enrichment with respect to solar) and depleted carbon (at least by an order of magnitude) and oxygen (a factor of about 1/4, see \\citealt{hub08}). In addition, \\tc\\ is also a binary system, with short period (2.2d, \\citealt{llo95}) and small eccentricity ($e$=0.13, \\citealt{hub08}). The companion remains undetected in the visible spectrum (its flux contributes to $<$0.1\\% of the total flux) and should therefore be of a much later spectral type ($M\\sim1$\\,M$_{\\odot}$, \\citealt{hub08}). The binarity and anomalous abundances might suggest that the blue straggler character of \\tc\\ results from a past episode of mass-transfer between these two stars.\\\\ Because of its peculiar properties, the spectrum of \\tc\\ was investigated several times, notably to search for the presence of a magnetic field \\citep{bor79,hub08}. The results are rather inconclusive, but an intriguing period of about 9 minutes was detected in the most recent spectropolarimetric results. \\\\ We decided to further study the star in the high-energy domain. As it was detected by Einstein and Rosat, \\tc\\ is known as the brightest X-ray source of IC\\,2602, but only approximate X-ray properties have up to now been derived. The current generation of X-ray facilities (Chandra, \\xmm) provides a detailed insight on the X-ray emission, especially thanks to their grating instruments. However, these observatories have observed only a handful of B stars at high spectral resolution ($\\tau$\\,Sco, $\\beta$\\,Cru, $\\epsilon$\\,Ori, Spica) and no blue straggler. The analysis of \\tc\\ thus nicely fills a gap, and could help better understand the high-energy characteristics of hot stars.\\\\ The paper is organized as follows: the data and reduction processes are presented in Sect. 2, the general properties of \\tc\\ in the X-ray domain are examined in Sect. 3.1, the detailed characteristics of the X-ray lines are derived in Sect. 3.2, and we finally conclude in Sect. 4. ", "conclusions": "\\xmm\\ observations revealed the atypical character of \\tc. Overall, its X-ray emission appears very soft as well as rather weak. Almost all the flux is found below 1\\,keV; indeed, spectral fits indicate a dominant temperature of about 0.2\\,keV. The total unabsorbed luminosity in the 0.1--10\\,keV range amounts to 1.0$\\times10^{31}$~erg\\,s$^{-1}$, which yields a $L_X/L_{BOL}$ ratio slightly lower than the `canonical' value for OB stars. Though no significant variability is detected during the 50\\,ks \\xmm\\ observation, an increase (resp. decrease) of the flux is clearly observed when comparing with previous Einstein (resp. ROSAT) observations. \\\\ The high-resolution spectrum of \\tc, revealed by the RGS, does not show any significant continuum emission but is solely composed of lines of H- and He-like ions of N, O, and Ne as well as some iron lines. To the resolution limits of the RGS, these lines appear narrow and unshifted: the observed width of these lines is mainly instrumental and the intrinsic width is limited to $<350$\\,\\kms. It must be noted that the nitrogen lines are anomalously strong: the spectral fits reveal a large enrichment in nitrogen (abundance 3 times solar), together with a strong deficit in carbon, which both agree well with the values found in the optical spectrum \\citep{hub08}. Neon, oxygen and iron also appear slightly depleted in the X-ray spectrum. For the {\\it fir} triplets, the forbidden component is fully suppressed while the intercombination lines are slightly stronger than the resonance component. The {\\it f/i} ratios suggests a formation radius rather close to the star, below 5 stellar radii.\\\\ With its soft X-ray spectrum and its narrow X-ray lines, \\tc\\ clearly appears different from O-type stars. Comparing \\tc\\ to other B-type objects with high-resolution spectra, we find that the temperature distribution is quite typical of `normal' B stars: the DEM is similar to that of $\\epsilon$\\,Ori and only slightly broader than for $\\beta$\\,Cru and Spica \\citep{zhe07}. Similar temperatures, fluxes and overall line properties were also found in a detailed analysis of $\\beta$\\,Cru \\citep{coh08}, though in the latter case, more precise constraints could be found on the formation radius. On the other hand, \\tc\\ seems very different from $\\tau$ Sco as far as softness, abundances, and flux level are concerned, but both objects present unshifted, narrow X-ray lines \\citep{mew,coh03}. Such lines are at odds with the wind-shock model, though it is still unclear how this model could apply in the case of low mass-loss rates, as those of B-type stars. Actually, narrow lines are often attributed to magnetic confinement. However, this process is still uncertain for \\tc, as magnetic field searches were inconclusive up to now \\citep{hub08}. As for $\\beta$\\,Cru, additional observations, especially polarimetric ones, are requested before \\tc\\ can be fully understood and the origin of its X-ray emission pinpointed." }, "0808/0808.3769_arXiv.txt": { "abstract": "The uniformity of the helium-to-hydrogen abundance ratio in X-ray emitting intracluster medium (ICM) is one of the commonly adopted assumptions in X-ray analyses of galaxy clusters and cosmological constraints derived from these measurements. In this work, we investigate the effect of He sedimentation on X-ray measurements of galaxy clusters in order to assess this assumption and associated systematic uncertainties. By solving a set of flow equations for a H-He plasma, we show that the helium-to-hydrogen mass ratio is significantly enhanced in the inner regions of clusters. The effect of He sedimentation, if not accounted for, introduces systematic biases in observable properties of clusters derived using X-ray observations. We show that these biases also introduce an apparent evolution in the observed gas mass fractions of X-ray luminous, dynamically relaxed clusters and hence biases in observational constraints on the dark energy equation of state parameter, $w$, derived from the cluster distance-redshift relation. The Hubble parameter derived from the combination of X-ray and Sunyaev-Zel'dovich effect (SZE) measurements is affected by the He sedimentation process as well. Future measurements aiming to constrain $w$ or $H_0$ to better than 10\\% may need to take into account the effect of He sedimentation. We propose that the evolution of gas mass fraction in the inner regions of clusters should provide unique observational diagnostics of the He sedimentation process. ", "introduction": "Clusters of galaxies are powerful cosmological probes and have the potential to constrain properties of dark energy and dark matter. Recent development in X-ray observations of galaxy clusters have produced a large statistical sample of clusters and start to deliver powerful cosmological constraints \\citep{Allen2004Constraints_on_,Allen2008Improved,Mantz2007New_constraints, Vikhlinin2008} that are complimentary to and competitive with other techniques (e.g., supernova, baryon acoustic oscillation, and weak lensing). This has motivated construction of the next-generation of X-ray satellite missions (e.g., \\emph{eROSITA}) to push the precision cosmological measurements based on large X-ray cluster surveys. However, in the era of precision cosmology, the use of clusters as sensitive cosmological probes require solid understanding of cluster gas physics, testing of simplifying assumptions, and assessing associated systematic uncertainties. One of the commonly adopted assumptions in X-ray cluster analyses include the uniformity of the helium-to-hydrogen abundance ratio with nearly primordial composition in X-ray emitting intracluster medium (ICM). At present, there is no observational test of this assumption, since both H and He in the ICM are fully ionized, which makes it difficult to measure their abundances using traditional spectroscopic techniques. Theoretically, on the other hand, it has long been suggested that heavier He nuclei slowly settle in the potential well of galaxy clusters and cause a concentration of He toward their center \\citep{Abramopoulos1981On_the_equilibr,Gilfanov1984Intracluster,Qin2000BARYON-DISTRIBU,Chuzhoy2003Gravitational,Chuzhoy2004Element,Ettori2006Effects}. In the era of precision cosmology, this could be a source of significant systematic uncertainties in X-ray measurements of galaxy clusters and cosmological parameter derived from these measurements \\citep{markevitch2007Helium_abundance}. Thus, the primary goal of the present work is to assess the validity of this assumption and associated systematic uncertainties in X-ray measurements of key cluster properties as well as cosmological parameters derived from these observations. In this work, we investigate the effects of He sedimentation on X-ray measurements of galaxy clusters by solving a set of diffusion equations for a H-He plasma in the ICM. By taking into account observed temperature profiles obtained by recent X-ray observations \\citep{Vikhlinin2006Chandra_Sample,2007Pratt, 2008Leccardi,2008George}, we show that the observed temperature drop in the cluster outskirts lead to a significant suppression of He sedimentation, compared to the results based on the isothermal cluster model \\citep{Chuzhoy2004Element}. Our analysis indicates that the He sedimentation has negligible effect on X-ray measurements in the outer regions of clusters (e.g., $r_{500}$), and it does not affect cluster mass measurements obtained at the sufficiently large cluster radius. The effect of He sedimentation, on the other hand, introduces increasingly larger biases in X-ray measurements in the inner regions and could affect cosmological constraints, including the dark energy equation of state parameter $w$ derived from distance-redshift relation as well as $H_0$ derived from the combination of X-ray and Sunyaev-Zel'dovich effect (SZE). The paper is organized as follows. In \\S~\\ref{sec:xray}, we describe the dependence of X-ray clusters measurements on He abundance in the ICM. The physics of He sedimentation and cluster models are discussed in \\S~\\ref{sec:diffusion}. In \\S~\\ref{sec:results}, we present results of our He sedimentation calculations and investigate their effects on cluster properties and cosmological constraints derived from X-ray cluster observations. Main conclusions are summarized in \\S~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} In this work, we investigate effects of He sedimentation on X-ray measurements of galaxy clusters and their implication for cosmological constraints derived from these observations. By solving a set of flow equations for a H-He plasma and using observationally motivated cluster models, we show that the efficiency of He sedimentation is significantly suppressed in the cluster outskirts due to the observed temperature drop, while it is dramatically enhanced in the cluster core regions. Our sedimentation model based on the observed temperature profile suggest that the effect of helium sedimentation is negligible at $r_{500}$, and it does not affect cluster mass measurements obtained at the sufficiently large cluster radius. However, the effect of sedimentation increases toward the inner regions of clusters and introduces biases in X-ray measurements of galaxy clusters. For example, at $r_{2500}$, biases in X-ray measurements of gas mass, total mass, and gas mass fractions, are at the level of $5-10\\%$. The effect of He sedimentation could also introduce biases in the estimate of the Hubble parameter derived from the combination of X-ray and SZE measurements, which could explain the observed offset in the X-ray+SZE derived $H_0$ and independent measurement from the Hubble Key project. We emphasize, however, that the magnitude of these biases depends sensitively on the cluster age, temperature, and magnetic and/or turbulent suppression in the ICM. We show that the process of He sedimentation introduces the apparent evolution in the observed gas mass fractions of X-ray luminous, dynamically relaxed clusters. The effect of He sedimentation could lead to biases in observational constraints of dark energy equation of state $w$ at a level of $\\lesssim$10\\%. These biases tend to make the value of $w$ more negative. Current measurements based on $f\\gas$ evolution \\citep{Allen2004Constraints_on_,Allen2008Improved} should not be significantly affected by these biases. However, future measurements aiming to constrain $w$ to better than 10\\% may need to take into account the effect of He sedimentation. For cosmological measurements, one way to minimize these biases is to extend the X-ray measurements to a radius well beyond $r_{2500}$. At the same time, the evolution of cluster gas mass fraction in the inner regions of clusters should provide unique observational diagnostics of the He sedimentation process in clusters." }, "0808/0808.1841_arXiv.txt": { "abstract": "The morphology of the outer rings of early-type spiral galaxies is compared to integrations of massless collisionless particles initially in nearly circular orbits. Particles are perturbed by a quadrupolar gravitational potential corresponding to a growing and secularly evolving bar. We find that outer rings with R1R2 morphology and pseudorings are exhibited by the simulations even though they lack gaseous dissipation. Simulations with stronger bars form pseudorings earlier and more quickly than those with weaker bars. We find that the R1 ring, perpendicular to the bar, is fragile and dissolves after a few bar rotation periods if the bar pattern speed increases by more than $\\sim 8\\%$, bar strength increases (by $\\gtrsim 140\\%$) after bar growth, or the bar is too strong ($Q_T>0.3$). If the bar slows down after formation, pseudoring morphology persists and the R2 ring perpendicular to the bar is populated due to resonance capture. The R2 ring remains misaligned with the bar and increases in ellipticity as the bar slows down. The R2 ring becomes scalloped and does not resemble any ringed galaxies if the bar slows down more than 3.5\\% suggesting that bars decrease in strength before they slow down this much. We compare the morphology of our simulations to B-band images of 9 ringed galaxies from the Ohio State University Bright Spiral Galaxy Survey, and we find a reasonable match in morphologies to R1R2' pseudorings seen within a few bar rotation periods of bar formation. Some of the features previously interpreted in terms of dissipative models may be due to transient structure associated with recent bar growth and evolution. ", "introduction": "Rings in barred galaxies can exist interior to the bar, encircling the bar or exterior to the bar. For a review on classification and properties of ringed galaxies see \\citet{buta96}. The outer rings of barred galaxies are classified as R1 or R2 depending upon whether the ring is oriented with major axis perpendicular to the bar (R1) or parallel to it (R2) (e.g., \\citealt{romero06}). If the ring is broken, partial or is a tightly wrapped spiral it is called a pseudoring and denoted R1' or R2'. Some galaxies contain both types of rings and are denoted R1R2' or R1R2. R1' and R2' morphologies were predicted as morphological patterns that would be expected near the outer Lindblad resonance (OLR) with the bar \\citep{schwarz81,schwarz84}. Rings are often the site of active star formation and so are prominent in blue visible band images, H$\\alpha$ narrow band images, and HI emission \\citep{buta96}. Orbital resonances, denoted Lindblad Resonances, occur at locations in the disk where \\begin{equation} \\Omega_b = \\Omega \\pm \\kappa/m \\end{equation} where $\\Omega_b$ is the angular rotation rate of the bar pattern and $m$ is an integer. Here $\\Omega(r)$ is the angular rotation rate of a star in a circular orbit at radius $r$ and $\\kappa(r)$ is the epicylic frequency. The $m=2$ OLR is that with $\\Omega_b = \\Omega + \\kappa/2$. Orbits of stars are often classified in terms of nearby periodic orbits that are closed in the frame rotating with the bar. Near resonances orbits become more elongated and have higher epicyclic amplitudes. Exterior to the OLR periodic orbits parallel to the bar are present whereas interior to the OLR both perpendicular and parallel periodic orbits are present. For a steady pattern, closed orbits interior to the OLR are expected to be aligned with major axis perpendicular to the bar whereas those exterior to the OLR are aligned parallel to it (e.g., \\citealt{cont89,kalnajs91}). A common assumption is that rings form because gas accumulates at resonances. This follows as gas clouds cannot follow self-intersecting orbits without colliding. Because of dissipation in the gas, the bar can exert a net torque on the gas leading to a transfer of angular momentum. The torque is expected to change sign at resonances so gas can move away from them or accumulate at them. The CR region is expected to be depopulated leading to gas concentrations at the OLR and ILR resonances. Gaseous rings form when gas collects into the largest periodic orbit near a resonance that does not cross another periodic orbit \\citep{schwarz84}. \\citet{schwarz81,schwarz84} first demonstrated the efficiency of this process. Other papers have confirmed and extended this work (e.g., \\citealt{combes85,byrd94,salo99,rau00,rau04}). Because dissipation is thought to be important, spiral like features and ovals that are not perfectly aligned with the bar, similar to those observed, are predicted. In some cases galaxy morphology and kinematics have not been successfully modeled with a single steady state bar component. Improvements in the models have been made with the addition of an additional exterior oval or spiral component (e.g., \\citealt{hunter88,lindblad96}). Previous work accounting for ring galaxy morphology has primarily simulated the gas dynamics using sticky particle simulations that incorporate dissipative or inelastic collisions. \\citet{rau00} ran N-body stellar simulations coupled with sticky gas particles. These simulations have self-consistent bars so that the orbits of the stars in the bars are consistent with the bar's gravitational potential. The disadvantage of using N-body simulations is that the properties of the bar such as its pattern speed and strength cannot be set. They can only be changed indirectly by varying the initial conditions of the simulations. An alternative approach is to set the bar perturbation strength, shape and pattern speed and search for likely bar parameters consistent with the properties of observed galaxies (e.g., \\citealt{salo99,rau04,rau08}). Previous work has explored the affect of bar strength and pattern speed on ring morphology (e.g., \\citealt{salo99,rau04,rau08}) and length of time since the bar grew (e.g., \\citealt{rau00,ann00}). Here we explore the role of bar evolution on ring galaxy morphology. By bar evolution we mean changes in bar pattern speed and strength during and after bar growth. N-body simulations lacking live halos predict long lived bars with nearly constant pattern speeds (e.g., \\citealt{voglis07}). However angular momentum transfer between a bar and the gas disk either interior or exterior to the bar or between a bar and a live halo can cause the pattern speed to vary (e.g., \\citealt{debattista98,bournaud02,das03,ath03,sellwood06,martinez06}). Thus constraints on the secular evolution of bars could tell us about the coupling between bars, gas and dark halos. Gas and stars exterior to a bar are sufficiently distant and moving sufficiently slowly compared to the bar that they are unlikely to cause strong perturbations on the orbits of stars in the bar. Because a calculation of the gravitational potential involves a convolution with an inverse square law function, high order Fourier components are felt only extremely weakly exterior to the bar. The dominant potential term exterior to the bar is the quadrupolar term which decreases with radius to the third power, $\\Phi \\propto r^{-3}$. Here we explore the role of a changing quadrupolar potential field on the morphology of stars exterior to a bar. In this work we focus on collisionless stellar orbits and leave investigating the study of dissipative effects for future study. In Section 2 we describe our simulations and present the results obtained by varying the parameters. In Section 3 we compare the results of our simulations with 9 galaxies from the Ohio State University Bright Spiral Galaxy Survey (\\citet{eskridge02}, hereafter OSUBSGS). Finally in Section 4 we summarize and discuss our results. ", "conclusions": "We have presented integrations of collisionless massless particles perturbed by growing and secularly evolving bar perturbations. We find that collisionless simulations can exhibit double ringed R1 and R2 outer ring morphology with rings both perpendicular (R1) and parallel (R2) to the bar. In the last period of bar growth, strong open spiral structure is exhibited resembling an R1' pseudoring. For 2-3 periods following bar growth R1 and R2 rings are seen with the R2 ring changing in orientation and azimuthal density contrast. Thus R1R2' pseudoring morphology is displayed within a few bar periods following bar growth. Our simulations start with particles in nearly circular orbits with velocity dispersions equivalent to 7 km/s for a 200 km/s rotation curve. This suggests sticky particle simulations have been successful in exhibiting R1R2 ring morphology because the velocity dispersion of orbits is damped and so particles are in initially nearly circular orbits. In our collisionless simulations we find that the outer rings with major axis perpendicular (R1) to the bar are fragile. If the bar pattern speed increases more than 8\\% after bar growth, or if the bar strength is higher than or increases past $|\\epsilon| \\gtrsim 0.16$ or $Q_T \\gtrsim 0.32$ the R1 outer ring will dissolve after $\\sim 20$ twenty bar periods. The simulations are then nearly mirror symmetric and do not display asymmetries typical of pseudorings. Stronger bars can form R1' pseudorings earlier. However if the bar strength $|\\epsilon| \\gtrsim 0.16$ or $Q_T \\gtrsim 0.32$ the R1 ring will dissolve after $\\sim 20$ bar rotation periods. If the bar strength increases to this value subsequent to formation, the R1 ring also dissolves. We find that a decrease in the bar pattern speed after bar growth causes particles to be captured in orbits parallel to the bar which are increased in epicyclic amplitude as the bar slows down. Strong R1 and elongated R2 rings persist in these simulations. Misalignments between the R2 ring and the bar also persist so the galaxy can exhibit R1R2' pseudoring morphology for a longer period of time. If the bar pattern speed slows down more than $\\sim 3.5\\%$ the R2 ring develops a scallop above and below the bar. As these are not observed in galaxies, bars probably do not slow down more than $\\sim 3.5\\%$ without also varying in strength. \\citet{sandage94} find that early type barred galaxies often have semi-detached outer rings (e.g, NGC 1543, \\citealt{buta96} and NGC 4457). These galaxies may contain bars that have increased in pattern speed or were once strong and so destroyed their R1 ring. If the bar weakens the R1 and R2 rings can be left behind as two nearly circular rings, similar to those observed in the unusual double outer ringed galaxy NGC 2273. We find that the morphology of our simulations resembles that of R1' ringed galaxies if the simulation time is chosen during or just after bar formation. We find we can match pseudoring morphology with simulations that have bar strengths estimated from the bar shapes. Stronger and longer spiral arms are seen later in the simulation and in more strongly barred systems. The constraint on simulation timescale suggests that R1' ring morphology is a signpost of recent bar formation. We note that sticky particle and SPH simulations exhibit R1 pseudoring morphology a few bar rotation periods longer than ours suggesting that the dissipationless simulations explored here underestimate the longevity of these features. We find that galaxies with R1R2' morphology are well matched by simulations a few bar rotation periods following bar growth. As R1 rings are fragile, we infer that these galaxies have had stable bars that have not experienced large changes in either pattern speed or strength. The exploration of parameter space in the collisionless dissipationless limit done here can be used by future work to differentiate between phenomena that would be exhibited by collisionless models and that that is a result of dissipation. A better understanding of the role of dissipation in affecting outer ring morphology should allow observationally based constraints on the secular evolution of bars. Only 10-20\\% of early type galaxies exhibit outer rings with pseudorings being more prevalent in later type galaxies \\citep{buta96}. Not all but most galaxies classified with outer rings are barred suggesting that only 15-40\\% of barred galaxies exhibit outer rings. Here we have found that R1' and R1R2' galaxies are likely to represent different times since bar formation with R1' galaxies representing an earlier timescale during or just after bar formation and R1R2 morphology representing galaxies with stable bars a few bar rotation periods following bar formation. Galaxies in these two transient categories probably comprise a significant fraction of all outer ring galaxies. This suggests that most outer ring galaxies represent morphology that is only present for a few bar rotation periods. It is interesting to ask what timescales these morphologies correspond to. Bar rotation periods for the ringed galaxies in our sample range from $\\sim 100-200$ Myr (see \\ref{tab:tab3}). The R1' classification, may only last a few bar rotations or 1/2 Gyr and the R1R2' classification only $\\sim$ 1 Gyr. Both of these timescales are short compared to the lifetime of a galaxy. Ringed galaxies lacking R1 rings may be longer lived but may provide evidence for bar evolution. It is likely that only a low fraction of barred galaxies might be considered systems that are not evolving secularly or have not formed in the last Gyr." }, "0808/0808.1593_arXiv.txt": { "abstract": "The nearby radio galaxy Centaurus A is poorly studied at high frequencies with conventional radio telescopes because of its very large angular size, but is one of a very few extragalactic objects to be detected and resolved by the {\\it Wilkinson Microwave Anisotropy Probe} ({\\it WMAP}). We have used the five-year {\\it WMAP} data for Cen~A to constrain the high-frequency radio spectra of the 10-degree giant lobes and to search for spectral changes as a function of position along the lobes. We show that the high-frequency radio spectra of the northern and southern giant lobes are significantly different: the spectrum of the southern lobe steepens monotonically (and is steeper further from the active nucleus) whereas the spectrum of the northern lobe remains consistent with a power law. The inferred differences in the northern and southern giant lobes may be the result of real differences in their high-energy particle acceleration histories, perhaps due to the influence of the northern middle lobe, an intermediate-scale feature which has no detectable southern counterpart. In light of these results, we discuss the prospects for {\\it Fermi Gamma-ray Space Telescope} detections of inverse-Compton emission from the giant lobes and the lobes' possible role in the production of the ultra-high energy cosmic rays (UHECR) detected by the Pierre Auger Observatory. We show that the possibility of a {\\it Fermi} detection depends sensitively on the physical conditions in the giant lobes, with the northern lobe more likely to be detected, and that any emission observed by {\\it Fermi} is likely to be dominated by photons at the soft end of the {\\it Fermi} energy band. On the other hand we argue that the estimated conditions in the giant lobes imply that UHECRs can be accelerated there, with a potentially detectable $\\gamma$-ray signature at TeV energies. ", "introduction": "\\label{intro} Centaurus A is the closest radio galaxy to us (we adopt $D = 3.7$ Mpc, the average of 5 distance indicators in Ferrarese \\etal\\ 2007). Its proximity makes it one of the brightest extragalactic radio sources in the sky at low frequencies (only exceeded by Cygnus A: Baars \\etal\\ 1977), but also means that the outer `giant' double lobes (throughout the paper we use the nomenclature adopted by Alvarez et al 2000) subtend an angle of $\\sim 10^\\circ$ on the sky, although their total physical size ($\\sim 600$ kpc in projection) is not unusually large for an radio galaxy of Cen A's luminosity. The large angular size of the lobes has prevented the type of {\\it spatially resolved}, multifrequency study of their spectral structure that is commonplace for more distant radio galaxies (e.g. Alexander \\& Leahy 1987). While detailed low-frequency maps of the giant lobes have been available for many years (e.g. Cooper, Price \\& Cole 1965), published radio data from ground-based observations only exist up to 5 GHz (Junkes \\etal\\ 1993) and the spectral study of Alvarez \\etal\\ (2000), involving (at many frequencies) painstaking graphical integration of contour maps, was only able to determine overall spectra for the giant lobes, finding no evidence for deviation from a single power law between 408 MHz and 5 GHz. The low-frequency two-point spectral index maps of Combi \\& Romero (1997), however, do show some evidence for position-dependent spectral steepening, particularly towards the end of the southern giant lobe. Cen~A is widely believed to be a restarting radio galaxy, in the sense that the inner lobes are the result of the current nuclear activity, while the giant outer lobes are the result of a previous outburst (Morganti \\etal\\ 1999). This picture is supported by the observation of hot thermal X-ray emission, apparently the result of strong shocks, surrounding both inner lobes (Kraft \\etal\\ 2003, 2007; Croston \\etal , in prep.) which implies that they are propagating supersonically into the intergalactic medium (IGM) of Cen~A and are disconnected from the giant lobes. However, the nature of the intermediate-scale northern middle lobe (NML: Morganti \\etal\\ 1999) is not completely clear in this model. In standard spectral ageing models (e.g. Jaffe \\& Perola 1973), we might expect to see spectral steepening at high frequencies in the giant lobes; a measurement of spectral ageing gives a model-dependent constraint on the time since the last injection of high-energy electrons into these lobes. Such a constraint on spectral age could be compared with other estimates of the dynamical age of the radio source, and would therefore be of considerable interest, but the work of Alvarez \\etal\\ (2000) only sets upper limits on this quantity, since they do not see any deviation from a power-law spectrum. \\begin{figure*} \\epsfxsize 17.5cm \\epsfbox{montage.eps} \\caption{Large-scale structure of Centaurus A with a resolution of $\\sim 0.83^\\circ$. Contours are at $1,2,4\\dots$ times the base level specified for each map, and take no account of background, except for the 408-MHz map, from which a constant background of 43 Jy beam$^{-1}$ has been subtracted. Top row, from left to right: 408 MHz (3 Jy beam$^{-1}$), 1.4 GHz (1 Jy beam$^{-1}$), 5 GHz (0.25 Jy beam$^{-1}$), 20 GHz (1.0 Jy beam$^{-1}$). Bottom row, from left to right: 30 GHz (0.5 Jy beam$^{-1}$), 40 GHz (1.0 Jy beam$^{-1}$), 60 GHz (1.0 Jy beam$^{-1}$), 90 GHz (2.0 Jy beam$^{-1}$). Circles in the bottom right-hand corner of each image indicate the beam size (diameter shows FWHM). The 20-GHz data have different noise characteristics from the other {\\it WMAP} images because they have not been convolved with a Gaussian; instrumental noise on scales smaller than the effective beam is therefore visible. See the text (Section \\ref{wmap-pro}) for discussion of the convolution, effective resolution and beam area of these images.} \\label{convolved-data} \\end{figure*} The giant lobes of Cen~A are also interesting because they are predicted to be strong sources of inverse-Compton emission as the relativistic electrons in the lobes scatter cosmic microwave background (CMB) photons to high energies; a detection of inverse-Compton emission from Cen~A would constrain the magnetic field strength in the lobes of Fanaroff \\& Riley (1974) class I (hereafter FRI) radio sources in general: we have little information on the magnetic field strengths in these low-power radio galaxies at present. However, X-ray emission from this process would be distributed on similar scales to the giant lobes, making it hard to detect. Cooke, Lawrence \\& Perola (1978) claimed an early detection of the giant lobes using {\\it Ariel V}, but Marshall \\& Clark (1981) argued that this was the result of point source contamination, placing a much lower upper limit on the flux from {\\it SAS 3} observations. At soft X-ray energies (e.g. Arp 1994) the situation is seriously confused by the presence of known thermal X-ray emission from the interstellar medium of the host galaxy, which more recently has been extensively studied with {\\it Chandra} and {\\it XMM-Newton} (e.g.\\ Kraft \\etal\\ 2003), and is also hard to study because the X-ray emission fills the field of view of modern soft X-ray imaging instruments such as {\\it ROSAT} (Arp 1994), {\\it XMM}, and {\\it ASCA} (Isobe \\etal\\ 2001), presenting almost insuperable problems of background modelling and subtraction. However, the spectrum of the inverse-Compton emission should be hard up to high energies (exactly how high depends on the model adopted for the electron energy spectrum, as we will discuss below). There are thus also interesting constraints from observations at MeV to GeV energies made with the {\\it Compton Gamma-Ray Observatory} (e.g. Steinle \\etal\\ 1998; Sreekumar \\etal\\ 1999), which do not detect the giant lobes but again set upper limits on their high-energy flux densities. More sensitive hard X-ray/$\\gamma$-ray observations exist with wide-field instruments like {\\it INTEGRAL} (e.g. Rothschild \\etal\\ 2006) and the {\\it Swift} Burst Alert Telescope (e.g. Markwardt \\etal\\ 2005) but as these are coded-aperture instruments they have limited sensitivity to extended emission (see e.g. Renaud \\etal\\ 2007). At present, therefore, there is no unambiguous detection of X-ray or $\\gamma$-ray inverse-Compton emission from the giant lobes of Cen~A. One of us has shown (Cheung 2007) that {\\it Fermi}\\footnote{Formerly known as the {\\it Gamma-ray Large Area Space Telescope}, {\\it GLAST}.} may have the sensitivity to detect inverse-Compton emission off the CMB from the lobes of Cen~A at energies from $\\sim$100 MeV to 10 GeV. However, the details of this depend on modelling of the electron energy spectrum at high energies, which in turn depends on high-frequency radio data. Finally, Cen~A's giant lobes are possible sources of ultra-high energy cosmic rays (UHECRs). Cen~A's proximity means that all aspects of the active galaxy -- central AGN, inner jets and lobes, and giant lobes -- have long been considered as possible UHECR accelerators (see e.g. Cavallo 1978; Romero \\etal\\ 1996, and, more recently, Gureev \\& Troitsky 2008 and references therein). Interest in Cen~A has been spurred by the remarkable discovery that 2 of the 27 UHECR events detected so far by the Pierre Auger Observatory (hereafter `PAO'; Abraham \\etal\\ 2007) appear to be arriving from the direction of the centre of Cen~A, while at least 2 additional events may be associated with it (e.g. Gorbunov \\etal\\ 2008a; Wibig \\& Wolfendale 2007; Fargion 2008) due to the large angular extent of the giant radio lobes (Gorbunov \\etal\\ 2008b; Moskalenko \\etal\\ 2008). Most scenarios discussed in the literature to date assume that UHECRs are produced near the supermassive black hole (SMBH) or in the inner jets (e.g. Cuoco \\& Hannestad 2008; Kachelriess, Ostapchenko \\& Tomas 2008), but an explanation in terms of the giant lobes has the advantage that it can easily account for the PAO events seen on larger scales. In order to investigate this quantitatively we need information about the magnetic field strengths and the {\\it leptonic} particle acceleration in these lobes, which can be provided by a combination of high-frequency radio observations and inverse-Compton constraints or measurements. The {\\it Wilkinson Microwave Anisotropy Probe} ({\\it WMAP}) has observed the whole sky at frequencies around 20, 30, 40, 60 and 90 GHz (known as K, Ka, Q, V and W bands respectively) with the aim of measuring structure in the CMB (e.g.\\ Hinshaw \\etal\\ 2008). The currently available {\\it WMAP} data represent 5 years of observations. Cen~A is clearly detected, and spatially resolved, in the {\\it WMAP} observations at all frequencies (e.g. Page \\etal\\ 2007) and Israel \\etal\\ (2008) have recently presented {\\it WMAP}-derived measurements of the flux density of the whole source, showing that there is clear steepening in the integrated spectrum at high frequencies. Thus the data are available to carry out a study of the variation of the radio spectrum as a function of position, to fit spectral ageing models to the large-scale lobes and investigate whether we can learn anything about the source dynamics, and to make predictions of the expected inverse-Compton emission from the giant lobes. In this paper we present the results of such a study. We first combine the 5-year {\\it WMAP} data on Cen~A with single-dish radio images at lower frequencies to make spatially resolved measurements of the radio spectra from 408 MHz to 90 GHz. We then discuss the implications of the high-frequency detections for the dynamics of Cen~A, for possible inverse-Compton detections of the giant lobes at high energies, and for acceleration of UHECRs and their possible $\\gamma$-ray emission signatures. ", "conclusions": "" }, "0808/0808.3305_arXiv.txt": { "abstract": "In this paper, we review some of the properties of dense molecular cloud cores. The results presented here rely on three-dimensional numerical simulations of isothermal, magnetized, turbulent, and self-gravitating molecular clouds (MCs) in which dense core form as a consequence of the gravo-turbulent fragmentation of the clouds. In particular we discuss issues related to the mass spectrum of the cores, their lifetimes and their virial balance. ", "introduction": "We performed 3D numerical simulations of magnetized, self-gravitating, and turbulent isothermal MCs using the TVD code (Kim et al. 1999) on grids with $256^{3}$ and $512^{3}$ cells (V\\'{a}zquez-Semadeni et al. 2005; Dib et al. 2007a; Dib et al. 2008a). The basic features of these simulations are: Turbulence is driven until it is fully developed (at least for 2 crossing timescales) before gravity is turned on. The Poisson equation is solved to account for the self-gravity of the gas using a standard Fourier algorithm. Turbulence is constantly driven in the simulation box following the algorithm of Stone et al. (1998). The kinetic energy input rate is adjusted such as to maintain a constant rms sonic Mach number $M_{s}=10$. Kinetic energy is injected at large scales, in the wave number range $k=1-2$. Periodic boundary conditions are used in the three directions. In physical units, the simulations have a linear size of 4 pc, an average number density of $\\bar{n}=500$ cm$^{-3}$, a temperature of 11.4 K, a sound speed of $0.2$ km s$^{-1}$, and an initial {\\it rms} velocity of 2 km s$^{-1}$. The Jeans number of the box is $J_{box}=4$ (number of Jeans masses in the box is $M_{box}/M_{J,box}=J_{box}^{3}=64$, where $M_{box}=1887$ M$_{\\odot}$). The simulations vary by the strength of the magnetic field in the box with $B_{0}= 0, 4.6, 14.5$, and 45.8 $\\mu$G for the non-magnetized, the strongly supercritical, the mildly supercritical, and the subcritical cloud models, respectively. Correspondingly, the plasma beta and mass-to-magnetic flux (normalized for the critical value for collapse $M/\\phi =(4 \\pi^{2} G)^{-1/2}$; Nakano \\& Nakamura 1978) values of the box are $\\beta_{p,box}=\\infty, 1, 0.1$, and $0.01$, and $\\mu_{box}=\\infty, 8.8, 2.8$, and $0.9$, respectively. Cores are identified using a clump-finding algorithm that is based on a density threshold criterion and a friend-of-friend approach as described in Dib et al. (2007a). We restrict our selection of cores to epochs where the Truelove criterion (Truelove et al. 1997) is not violated in any of them. Thus, the derived properties of our numerical cores can be best compared to those of starless prestellar cores. ", "conclusions": "" }, "0808/0808.3902_arXiv.txt": { "abstract": "We discuss the bounds on the mass of Dark Matter (DM) particles, coming from the analysis of DM phase-space distribution in dwarf spheroidal galaxies (dSphs). After reviewing the existing approaches, we choose two methods to derive such a bound. The first one depends on the information about the current phase space distribution of DM particles only, while the second one uses both the initial and final distributions. We discuss the recent data on dSphs as well as astronomical uncertainties in relevant parameters. As an application, we present lower bounds on the mass of DM particles, coming from various dSphs, using both methods. The model-independent bound holds for any type of fermionic DM. Stronger, model-dependent bounds are quoted for several DM models (thermal relics, non-resonantly and resonantly produced sterile neutrinos, etc.). The latter bounds rely on the assumption that baryonic feedback cannot significantly increase the maximum of a distribution function of DM particles. For the scenario in which all the DM is made of sterile neutrinos produced via non-resonant mixing with the active neutrinos (NRP) this gives $m_{\\dw}>1.7$ keV. Combining these results in their most conservative form with the X-ray bounds of DM decay lines, we conclude that the NRP scenario remains allowed in a very narrow parameter window only. This conclusion is independent of the results of the Lyman-alpha analysis. The DM model in which sterile neutrinos are resonantly produced in the presence of lepton asymmetry remains viable. Within the minimal neutrino extension of the Standard Model (the $\\nu$MSM), both mass and the mixing angle of the DM sterile neutrino are bounded from above and below, which suggests the possibility for its experimental search. ", "introduction": "\\label{sec:introduction} The nature of Dark Matter is one of the most intriguing questions of particle astrophysics. Its resolution would have a profound impact on the development of particle physics beyond the Standard Model. Although the possibility of having massive compact halo objects (MACHOs) as a dominant form of DM is still under debate (see recent discussion in~\\cite{Calchi:07} and references therein), it is widely believed that Dark Matter is composed of non-baryonic particles. However, the Standard Model of elementary particles does not contain a viable Dark Matter particle candidate -- a massive, neutral and long-lived particle. Active neutrinos, which are both neutral and stable, form structures in a top-down fashion~\\cite{Zeldovich:70,Bisnovatyi:80,Bond:80,Doroshkevich:81,Bond:83}, and thus cannot produce the observed quantity of early-type galaxies~\\cite[see e.g.][]{White:83,Peebles:84a}. Therefore, the DM particle hypothesis implies the extension of the Standard Model (SM). The DM particle candidates may have very different masses (for reviews of DM candidates see e.g.~\\cite{Bergstrom:00,Bertone:05,Carr:06,Taoso:07}): massive gravitons with the mass $\\sim 10^{-19}\\ev$~\\cite{Dubovsky:04}, axions with the mass $\\sim 10^{-6}\\ev$~\\cite{Holman:83}, sterile neutrinos having mass in the keV range~\\cite{Dodelson:93}, sypersymmetric (SUSY) particles (gravitinos~\\cite{Pagels:82}, neutralinos~\\cite{Haber:85}, axinos~\\cite{Covi:99} with their masses ranging from eV to hundreds GeV, supersymmetric Q-balls~\\cite{Kusenko:97b}, WIMPZILLAs with the mass $\\sim 10^{13}\\gev$~\\cite{Kuzmin:98,Chung:99}, and many others). Thus, the mass of DM particles becomes an important characteristic which may help to distinguish between various DM candidates and, more importantly, may help to differentiate among different models beyond the SM. It was suggested in~\\cite{Tremaine:79} that quite a robust and model-independent \\emph{lower bound} on the mass of DM particles can be obtained by considering phase space density evolution of compact astrophysical objects, most notably dwarf spheroidal satellites (dSphs) of the Milky Ways. The idea was developed further in a number of works (see e.g.~\\cite{Madsen:84,Madsen:91,Madsen:90,Dalcanton:00,Hogan:00,Madsen:00}).% Another way to distinguish between various DM models, and particularly to put a bound on the DM mass, is the analysis of the Lyman-$\\alpha$ (Ly-$\\alpha$) forest data~\\cite{Hui:97,Gnedin:01,Weinberg:03}.\\footnote{Absorption feature by neutral hydrogen at $\\lambda = 1216$~\\AA~at different redshifts in the spectra of distant quasars.} % This method essentially constrains the possible shape of the power spectrum of density fluctuations at comoving scales $\\sim$Mpc. Assuming \\emph{a DM model}, (i.e. a particular primordial velocity distribution of DM particles), one can obtain a relationship between the DM particle mass \\emph{in this model} and the shape of the power spectrum, probed by Ly-$\\alpha$. Although very promising, the Ly-$\\alpha$ method is very complicated and indirect. First of all, under the assumption that the distribution of the neutral hydrogen traces that of the DM, one can reconstruct the power spectrum of density fluctuations at redshifts $z \\sim 2-5$ from the statistics of Lyman-$\\alpha$ absorption lines. One can then perform a fit of the Lyman-$\\alpha$ data (often together with the measurements of anisotropy of temperature of cosmic microwave background and the data of large-scale structure surveys), to extract the information about cosmological models. This is usually done by using the Monte-Carlo Markov chain technique~\\cite{Lewis:02}. At redshifts probed by Ly-$\\alpha$, the evolution of structure has already entered the (mildly) non-linear stage. Therefore, to properly relate the measured power spectrum with the parameters of a given cosmological model one would have to perform a prohibitively large number of hydrodynamic numerical simulations. Therefore, various simplifying approximations have to be realized~\\cite{Theuns:98,Gnedin:01,McDonald:05,Viel:04,Viel:05,Viel:05b,Viel:05c,Regan:06a}. Apart from these computational difficulties, the physics entering the \\lya analysis is complicated, and not yet fully understood~(see e.g.~\\cite{Kim:07a,Bolton:07a,Viel:2002ui,Viel:2001hd,Viel:2003fx}). Moreover, the DM particles can significantly influence the background physics, further complicating the Ly-$\\alpha$ analysis~\\cite{Biermann:06,Gao:07,Stasielak:07}. For a recent overview of the \\lya method see e.g.~\\cite{Boyarsky:08c}. The systematic uncertainties associated with both computational difficulties and complicated physics of \\lya systems are not fully explored. Therefore, \\emph{it is very important to have a lower mass bound on DM particles from more direct and simple considerations}. In this paper we discuss the Tremaine-Gunn and related DM mass bounds based on the phase-space density considerations as well as possible ways to strengthen them for several DM models. The obtained phase-space density bounds are weaker yet comparable with Ly-$\\alpha$ bounds and therefore provide an interesting alternative. We consider a class of the so-called ``generic'' DM models, where DM particles are produced thermally and decouple while being relativistic, thus having the (relativistic) Fermi-Dirac momentum spectrum. We also consider models of non-thermal DM production. In this case the primordial velocity spectrum of DM particles depends on the details of the production mechanism. We analyze the case when the velocity spectrum can be approximated by the \\emph{rescaled} Fermi-Dirac spectrum, or has two such components (a colder and a warmer one). A very important example of such a DM particle is the \\emph{sterile} (right-handed) neutrino. Although known as a DM candidate for some 15 years~\\cite{Dodelson:93}, recently sterile neutrinos have attracted a lot of attention. It was shown~\\cite{Asaka:05a} that if one adds three right-handed (sterile) neutrinos to the Standard Model, it is possible to explain simultaneously the data on neutrino oscillations (see e.g.~\\cite{Fogli:05,Strumia:06,Giunti:06} for a review) and the Dark Matter in the Universe, without introducing any new physics \\emph{above electro-weak scale} $M_W \\sim 100\\gev$. Moreover, if the masses of two of these particles are between $\\sim 100 \\MeV$ and electro-weak scale and are almost degenerate, it is also possible~\\cite{Asaka:2005pn} to generate the correct baryon asymmetry of the Universe (see e.g.~\\cite{Dolgov:97,Riotto:98}). The third (lightest) sterile neutrino can have mass in keV-MeV range\\footnote{There are several interesting astrophysical applications of $\\keV$ sterile neutrinos (see e.g. \\cite{Sommer:99,Kusenko:06a,Biermann:06,Hidaka:06,Hidaka:07,Stasielak:06} and references therein).} % and be coupled to the rest of the matter weakly enough to provide a viable (\\emph{cold} or \\emph{warm)} DM candidate. This theory, explaining the three observed phenomena ``beyond the SM'' within one consistent framework, is called the \\emph{$\\nu$MSM}~\\cite{Asaka:2005pn,Asaka:05a} (see also~\\cite{Shaposhnikov:07b}). Although weakly coupled, the DM sterile neutrino in the $\\nu$MSM can be produced in the correct quanities to account for all of the DM. There are several mechanisms of production: non-resonant active-sterile neutrino oscillations (\\textbf{\\emph{non-resonant production}} mechanism, \\textbf{NRP}) ~\\cite{Dodelson:93,Dolgov:00,Abazajian:01a,Asaka:06b,Asaka:06c}, resonant active-sterile neutrino oscillations in the presence of lepton asymmetry (\\textbf{\\emph{resonant production}} mechanism, \\textbf{RP})~\\cite{Shi:98,Abazajian:01a,Shaposhnikov:08a,Laine:08a}, decay of the gauge-singlet scalar field~\\cite{Shaposhnikov:06} (see also \\cite{Kusenko:06a,Petraki:07,Petraki:08}). The \\lya analysis of the sterile neutrino DM, produced via NRP scenario was performed in a number of works~\\cite{Viel:06,Seljak:06,Viel:08}. These bounds were recently revisited in~\\cite{Boyarsky:08c}, using the SDSS \\lya dataset together with WMAP5~\\cite{Dunkley:2008ie}. The lower bound on the DM mass was found to be in this case $8\\kev$ (at $99.7\\%$~CL). Ref.~\\cite{Boyarsky:08c} also analyzed a more general case CWDM case : a mixture of NRP sterile neutrino with cold DM (see also~\\cite{Palazzo:07}). These results were applied to the RP produced sterile neutrino in~\\cite{Boyarsky:08d}. It was shown that the mass as low as $2\\kev$ is compatible with \\lya data. In this paper we will analyze in detail restrictions on the sterile neutrinos, produced via first two production mechanisms and briefly comment on the third one in the Discussion. In the case of non-resonant production the primordial velocity spectrum is approximately \\emph{proportional} to the Fermi-Dirac distribution~\\cite{Dolgov:00,Hansen:01} (the exact spectrum was calculated in ~\\cite{Asaka:06b,Asaka:06c}). The paper is organized as follows. In Section~\\ref{sec:overview} we review DM mass bounds, based on the phase-space density arguments. In Section~\\ref{sec:maxim-coarse-grain} we introduce the concept of maximal coarse-graining and propose a conservative modification of the original Tremaine-Gunn bound. In Section~\\ref{sec:analys-meas-valu} we analyze new observational data on recently discovered dSphs (see~\\cite{Gilmore:07,Simon:07} and references therein), and use it to determine the phase-space density of these objects. Special attention is paid to determine various systematic uncertainties of measured values. Our results are summarized in Section~\\ref{sec:bounds}. We conclude with the discussion of the results, analysis of possible uncertainties and outlook for the further improvement of the mass bounds in Section~\\ref{sec:discussion}. ", "conclusions": "\\label{sec:discussion} In this paper we suggested that a conservative way to put the bound on a DM particle mass may be based on the requirements that the maximum of the observed coarse-grained phase space density should not exceed the maximum of the initial distribution function of the DM particles. The maximum of the coarse-graned distribution function in the final state may be conservatively estimated from the observed quantities. This bound relies on the assumption that the maximum of the distribution function was not significantly increased by the interaction with baryons. Although DM consists of the non-interacting particles, the remaining part of the galaxy -- the baryons -- interact with one another and dissipate their energy, finally concentrating towards the center. The baryons, which are condensed in the center, influence the shape of DM halo gravitationally, increasing the central DM density~\\cite{Blumenthal:86,Gnedin:04}. The opposite effect is the energy feedback from SNae, galactic winds and reionization, which creates the strong outflow, significantly decreasing the mass of the gas and thereby affecting the DM halo shape. Such a feedback is thought to be responsible to the formation of dwarf spheroidals from gas-rich dwarf spiral/irregular galaxies~\\cite{Lin:83,Moore:94,Mastropietro:05,Mayer:06,Mayer:07}. Clearly both gas condensation and feedback strongly influence the central PSD of DM~\\cite{Read:06}, and in principle can lead to the violation of the inequality (\\ref{ff}). Numerical studies of galaxy mergers show that baryons can lead to the increase of the phase-space density during the merger (see e.g.~\\cite{Naab:07a}). However, the method used in this work -- coarse-graining of the PSD over a large phase-space region -- reduces the influence of baryons. Indeed, we take the spatial averaging over the radius $R \\sim r_h$, which includes external part of the system, where the amount of baryons is small. Additional studies are necessary to estimate effects of baryons and make our bounds more robust. We plan to address these issues elsewhere. We would also like to stress that the initial velocities of DM particles in our approach are \\emph{thermal} velocities and they should not be confused with the so-called \\emph{Zeldovich} velocities~\\cite{Zeldovich:70}. Numerical simulations of galaxy formation do not start at the time, when the DM phase-space distribution is spatially uniform (redshifts $z\\gtrsim 10^3$). Instead, the initial (linear) stage of the structure formation is computed analytically in the framework of the so-called \\emph{Zeldovich approximation}~\\cite{Zeldovich:70}. This approximation is commonly used to set up initial conditions for the numerical simulations of non-linear stage of structure formation~\\cite{Bertschinger:95,Klypin:97,Klypin:00}, which start at redshifts $z\\sim 10$. The peculiar (\\emph{Zeldovich}) velocities acquired by DM particles at this stage due to structure formation and included into the initial conditions are normally $\\sigma \\sim 10\\km/\\sec$. Apart from Zeldovich velocities, DM particles also possess thermal velocities, which are discussed in this paper. For cold enough Dark Matter these thermal velocities are much smaller than Zeldovich ones and, thus, are often neglected and not included into initial conditions. Therefore, the numerical studies of PSD evolution\\footnote{Most of these studies use the quantity $Q(r) = \\rho(r)/\\sigma^3(r)$ as a PSD estimator} (see e.g.~\\cite{Taylor:01,Peirani:06,Peirani:07,Romano-Diaz:07,Hoffman:07,Romano-Diaz:06}) essentially investigate the change of PSD from Zeldovich to final stage. It was found in some of these works that the PSD changes by $10^2-10^3$ in the process of collapse~\\cite{Peirani:06}. This change of PSD can be understood as being simply an evolution from initial Zeldovich velocities $\\sigma_i \\sim 10\\km/\\sec$ to the final (virial) ones $\\sigma_f \\sim 10^2\\km/\\sec$ (with $Q_i/Q_f \\sim (\\sigma_f/\\sigma_i)^3 \\sim 10^3$). Because initial thermal velocities may be much smaller than Zeldovich ones, initial PSD may differ from the final (observed) PSD not by 2--3, but by many orders of magnitude. This fact does not contradict to the results of simulations, described in e.g.~\\cite{Peirani:06} and, therefore, cannot be used to obtain an \\emph{upper} bound on the mass of DM particles (c.f~\\cite{Boyanovsky:08,Gorbunov:08a}). This work was mostly concentrated on restrictions on the mass of the sterile neutrino DM, produced in through the non-resonant oscillations with active neutrino (NRP scenario). We see that our results (Section~\\ref{sec:bounds}) strongly disfavor such sterile neutrinos as the single DM component. This conclusion is not based on the \\lya method and therefore is not subject to its uncertainties (discussed in the Introduction). However, several uncertainties can affect this conclusion, the major being baryonic feedback. To make this result really robust, apart from further modeling of the baryonic influence, one needs to strengthen the tension between upper and lower mass bounds discussed in this paper. This is plausible and may be done either by improving the X-ray bounds with new observations or by strengthening the PSD consideration, which is in the first place related to better measurements of kinematics of dSphs. In the presence of lepton asymmetry, the resonant production (RP) of sterile neutrino DM takes place~\\cite{Shi:98}. This mechanism is more efficient~\\cite{Shi:98,Laine:08a,Shaposhnikov:08a} than the NRP scenario and allows to achieve required DM abundance for weaker mixings (c.f. Fig.~4 in~\\cite{Laine:08a}). This lifts the upper bound on the DM particle mass in this scenario up to $\\sim 50$~keV. At the same time, for the same mass the primordial velocity distribution of RP sterile neutrino DM is colder than in NRP one. This $f_{max}$ is as much as the order of magnitude bigger than~(\\ref{eq:16}) (c.f.~\\cite{Laine:08a}). This brings down by a factor $\\sim 2$ the analog of the mass bound~(\\ref{eq:29}). Analyzing available spectra for a range of lepton asymmetries, we see that models with $m_\\sf \\gtrsim 1 \\kev$ are allowed. Thus, there is a large open ``window'' of allowed DM masses (c.f. Fig.\\ref{fig:sf-window}). However, as the dependence of the velocity spectrum on the lepton assymetry is not monotonic, to obtain the exact shape of the lower bound on the mass at given mixing angle more work is needed. Nevertheless, our results show that the sterile neutrinos, produced in the presence of lepton asymmetry, are viable DM candidates, allowed by all current bounds. {Finally, we would like to comment on the mechanism of production of sterile neutrinos from decay of massive scalar field, for example the inflaton~\\cite{Shaposhnikov:06} (for other models see~\\cite{Kusenko:06a,Petraki:07,Petraki:08,Gorbunov:08a}). The primordial phase-space distribution function for this case was computed e.g. in~\\cite{Shaposhnikov:06,Petraki:07,Boyanovsky:08b,Gorbunov:08a}. Maximal value of phase-space density for this distribution is that of degenerate Fermi gas. Notice that the distribution functions in~\\cite{Shaposhnikov:06,Petraki:07,Gorbunov:08a} $f(p)$ is formally unbounded for small momenta: $f(p) \\sim p^{-1/2}$. From this one can easily find that the fraction of particles, having maximal phase-space density, is $\\sim 10^{-8}$. As only this small fraction of all particles has maximal phase-space density, we expect the mass bound in this case to be stronger than~(\\ref{eq:28}). The detailed analysis will be presented elsewhere. } \\bigskip After this work has been completed, we received a draft of the paper~\\cite{Gorbunov:08b}, where similar issues have been considered. Our results are consistent with those of discussed in~\\cite{Gorbunov:08b} wherever they overlap." }, "0808/0808.3419_arXiv.txt": { "abstract": "The equation of state (EoS) of dark energy $w$ remains elusive despite enormous experimental efforts to pin down its value and its time variation. Yet it is the single most important handle we have in our understanding of one of the most mysterious puzzle in nature, dark energy. This letter proposes a new method for measuring the EoS of dark energy by using the gravitational waves (GW) of black hole binaries. The method described here offers an alternative to the standard way of large scale surveys. It is well known that the mass of a black hole changes due to the accretion of dark energy but at an extremely slow rate. However, a binary of supermassive black holes (SBH) radiates gravitational waves with a power proportional to the masses of these accreting stars and thereby carries information on dark energy. These waves can propagate through the vastness of structure in the universe unimpeded. The orbital changes of the binary, induced by the energy loss from gravitational radiation, receive a large contribution from dark energy accretion. This contribution is directly proportional to $(1+w)$ and is dominant for SBH binaries with separation $R \\ge 1000$ parsec, thereby accelerating the merging process for $w > -1$ or ripping the stars apart for phantom dark energy with $w < -1$. Such orbital changes, therefore $w$, can be detected with LIGO and LISA near merging time, or with X-ray and radio measurements of Chandra and VLBA experiments. ", "introduction": "\\label{sec:intro} One of our most crucial questions about nature at present is: what is dark energy? The fact that our universe is accelerating \\cite{wmap,lauramelchiorri} and that dark energy constitutes about $70 percent$ of the total energy density of the universe, are well established by now. Many theoretical models have been put forth which cast dark energy in the form of a cosmic fluid \\cite{quintessence,kessence,transplanck} with time variations in its equation of state $w(z)$. Yet the simplest explanation for dark energy remains to be a pure cosmological constant (cc) $\\Lambda$. The trouble we face in understanding dark energy does not stem from a shortage of dark energy models, with $w(z_0) \\simeq -1$, that mimick at present the behavour of $\\Lambda$ and give rise to the observed acceleration of the universe. The puzzle rather lies on identifying which one of these possible candidates is the correct one. The best way to discriminate among the various possibilites and a pure $cc$, $\\Lambda$, is to experimentally measure the time variations of the dark energy equation of state $w(z)$. So far a popular parametrization for $w(z)$ is the linear one $w(z)=w_0 + w_{1} z +...$ with $w_1 = 0$ for a pure {\\it cc} \\cite{eos}. A knowledge of $w(z)$ is crucial for not only understanding the present accelerated expansion of the universe but also for making predictions for its future evolution and destiny. If $w(z) \\ll -1$ then dark energy is a phantom \\cite{phantom} which leads the universe to a Big Rip in the future. If $w = -1$ then we are probably \\cite{laurads} facing an eternal DeSitter state \\cite{eternalds} which at least at the classical level implies constant temperature and entropy therefore a cosmic heat death \\cite{fredlaura, fred}. Other forms of $w(z)$ can also allow for a Big Crunch \\cite{crunch} or bounces \\cite{bounce}. At present we can not infer which destiny our universe will meet without a better knowldege of $w(z)$. Major experimental efforts for pinning down the value and time-variations of $w(z)$ are under way through large scale surveys from CMB \\cite{wmap}, large scale structure \\cite{SDSS}, and SN1a observations. The endeavor of measuring, to confident precision, such small time variations in $w(z)$ has proven extremely difficult, partly due to the inherited errors in the experiments that are not instrumental but which originate from noise accumulated from the background and foreground effects through which the signal we receive has propagated. In order to minimize such errors associated with the propagation of the signal through the vastness of structure in the universe, we would like to propose in this letter a complimentary method for measuring $w(z)$. This method uses compact and localized objects, such as Black Holes, for acquiring information about $w(z)$, by exploiting the gravitational waves these objects emit when they are in binaries. The advantage of this method is twofold: first, gravitational waves propagate undisturbed through structure; and second, we have existing experiments which are either already operational or will be in the near future, such as LIGO and LISA missions designed to detect these binaries gravitational waves, or Chandra and VLBA experiments designed for X-ray and radio measurements. The accretion of dark energy by Black Holes is reviewed in Sec.2., including a review of the main parameters of binaries and gravitational waves, useful for our purposed. Sec.3presents the method we propose, along with an investigation, discussion and some illustrations, on how gravitational waves from SBH's binaries can be used for extrapolating the equation of state $w(z)$ of dark energy. ", "conclusions": "" }, "0808/0808.3713_arXiv.txt": { "abstract": "{Characterising the circumstellar dust around nearby main sequence stars is a necessary step in understanding the planetary formation process and is crucial for future life-finding space missions such as ESA's \\darwin{} or NASA's Terrestrial Planet Finder (TPF). Besides paving the technological way to \\darwin/TPF, the space-based infrared interferometers \\peg{} and FKSI (Fourier-Kelvin Stellar Interferometer) will be valuable scientific precursors.} {We investigate the performance of \\peg{} and FKSI for exozodiacal disc detection and compare the results with ground-based nulling interferometers.} {We used the GENIEsim software (Absil et al. 2006) which was designed and validated to study the performance of ground-based nulling interferometers. The software has been adapted to simulate the performance of space-based nulling interferometers by disabling all atmospheric effects and by thoroughly implementing the perturbations induced by payload vibrations in the ambient space environment.} {Despite using relatively small telescopes ($\\leq$ 0.5\\,m), \\peg{} and FKSI are very efficient for exozodiacal disc detection. They are capable of detecting exozodiacal discs respectively 5 and 1 time as dense as the solar zodiacal cloud and they outperform any ground-based instrument. Unlike \\peg, FKSI can achieve this sensitivity for most targets of the \\darwin/TPF catalogue thanks to an appropriate combination of baseline length and observing wavelength. The sensitivity of \\peg{} could, however, be significantly boosted by considering a shorter interferometric baseline length.} {Besides their main scientific goal (characterising hot giant extrasolar planets), the space-based nulling interferometers \\peg{} and FKSI will be very efficient in assessing within a few minutes the level of circumstellar dust in the habitable zone around nearby main sequence stars down to the density of the solar zodiacal cloud. These space-based interferometers would be complementary to Antarctica-based instruments in terms of sky coverage and would be ideal instruments for preparing future life-finding space missions.} ", "introduction": "\\begin{figure*}[t] \\centering \\includegraphics[width=7cm]{pegase_CV3.eps}\\hspace{1.5 cm} \\includegraphics[width=6cm]{FKSI.eps} \\caption{Left: overview of the \\peg{} space-based interferometer. Two 0.4-m siderostats are flying in a linear configuration with the beam combiner spacecraft located in the middle of the formation. Right: representation of FKSI, showing the two 0.5-m siderostats located on a 12.5-m boom.} \\label{fig:pegase_FKSI} \\end{figure*} Nulling interferometry is the core technique of future life-finding space missions such as ESA's \\darwin{} \\citep{Fridlund:2006} and NASA's Terrestrial Planet Finder Interferometer \\citep[TPF-I,][]{Beichman:2006}. Observing in the mid-infrared (6-20\\,$\\mu$m), these missions would enable the spectroscopic characterisation of the atmosphere of habitable extrasolar planets orbiting nearby main sequence stars. This ability to study habitable distant planets strongly depends on the density of exozodiacal dust in the inner part of circumstellar discs, where the planets are supposed to be located. In particular, the detection of habitable terrestrial planets would be seriously hampered for stars presenting warm ($\\sim$300\\,K) exozodiacal dust more than 10 to 100 times as dense as our solar zodiacal disc, depending on stellar type, stellar distance and telescope diameter \\citep{beichman:2006b,defrere:2008}. Assessing the level of circumstellar dust around nearby main sequence stars is therefore a necessary pre-requisite for preparing the observing programme of \\darwin/TPF by reducing the risk of wasting time on sources for which exozodiacal light prevents Earth-like planet detection. In addition, the existence of planets is intrinsically linked to circumstellar discs and observing them provides an efficient way to study the formation, evolution and dynamics of planetary systems. At young ages, essentially all stars are surrounded by protoplanetary discs in which the planetary systems are believed to form \\citep{Meyer:2008}. In particular, the detection of gaps in these protoplanetary discs is very important for understanding the early dynamics of planets, including migration and orbital interaction. At older ages, photometric surveys primarily with IRAS, ISO, and Spitzer have revealed the presence of micron-sized grains around a large number of main sequence stars \\citep[see e.g.,][]{Trilling:2008,Hillebrand:2008}. This is interpreted as the sign of planetary activity, as the production of grains is believed (by analogy with the zodiacal cloud in our solar system) to be sustained by asteroid collisions and outgassing of comets in the first tens of astronomical units (AU). However, the presence of warm dust can generally not be unequivocally determined because the typical accuracy on both near-infrared photometric measurements and photospheric flux estimations is a few percent at best, limiting the sensitivity to typically 1000 times the density of our solar zodiacal cloud \\citep{Beichman:2006c}. Photometric measurements are therefore generally not sufficient to probe the innermost regions of the discs and interferometry is required to separate the starlight from the disc emission. Good examples are given by the detection of hot dust ($\\sim$1500\\,K) around Vega and $\\tau$ Cet with near-infrared interferometry at the CHARA array \\citep{Absil:2006b,Difolco:2007}. Nulling interferometry is a quite new technique even though it was initially proposed in 1978 \\citep{Bracewell:1978}. Several scientific observations using this technique have recently been carried out with the Bracewell Infrared Nulling Cryostat (BLINC, \\citealt{Hinz:2000}) instrument at the Multi-Mirror Telescope (MMT, Mont Hopkins, Arizona), with the Keck Interferometer Nuller (KIN, Hawaii, \\citealt{Serabyn:2006,Barry:2008,Serabyn:2008}), and are foreseen to begin in 2010 at the Large Binocular Telescope (Mount Graham, Arizona,\\citealt{Hinz:2008}). In Europe, ESA has initiated the study of a ground-based demonstrator for \\darwin, the Ground-based European Nulling Interferometer Experiment \\citep[GENIE,][]{gondoin:2004}. GENIE is a nulling interferometer conceived as a focal instrument for the VLTI which has been studied by ESA at the phase A level. Another European project is ALADDIN (Antarctic L-band Astrophysics Discovery Demonstrator for Interferometric Nulling, \\citealt{Foresto:2006}), a nulling interferometer project for Dome C, on the high Antarctic plateau. The performance of GENIE has been studied in detail (\\citealt{Absil:2006_paper1},\\citealt{Wallner:2006}) and recently compared to that of ALADDIN \\citep{Absil:2007_paper2}. Using 1-m collectors, ALADDIN would have an improved sensitivity with respect to GENIE working on 8-m telescopes, provided that it is placed above the turbulence boundary layer (about 30\\,m at Dome C). Circumstellar discs 30 times as dense as our local zodiacal cloud could be detected by ALADDIN around typical \\darwin/TPF targets in an integration time of few hours. The low atmospheric turbulence on the high Antarctic plateau is a significant advantage with respect to other astronomical sites and one of the main reasons for the very good sensitivity of ALADDIN. However, as for any other ground-based site, the atmosphere effects (turbulence and thermal background) are still major limitations to the performance and active compensation by real-time control systems are mandatory. Observing from space would provide an efficient solution to improve the sensitivity by getting rid of the harmful effect of the atmosphere. Two infrared nulling interferometers could achieve the detection of circumstellar dust discs from space (see Fig.~\\ref{fig:pegase_FKSI}): \\peg, a two-telescope interferometer based on three free-flying spacecraft \\citep{Leduigou:2006} and the Fourier-Kelvin Stellar interferometer (FKSI), a structurally-connected interferometer also composed of two telescopes \\citep{Danchi:2006}. These two missions have been initially designed to study hot extrasolar giant planets at high angular resolution in the near- to mid-infrared regime (respectively 1.5-6.0 $\\mu$m and 3.0-8.0 $\\mu$m). Besides their main scientific goal, they could also be particularly well suited for the detection of warm circumstellar dust in the habitable zone around nearby main sequence stars. The objective would be to provide a statistically significant survey of the amount of exozodiacal light in the habitable zone around the \\darwin/TPF targets, and its prevalence as a function of other stellar characteristics (age, spectral type, metallicity, presence of a cold debris disc, etc). Following our performance studies of ground-based instruments such as GENIE at Cerro Paranal (\\citealt{Absil:2006_paper1}, hereafter Paper\\,I) or ALADDIN on the high Antarctic plateau (\\citealt{Absil:2007_paper2}, hereafter Paper\\,II), the present study addresses the performance of space-based nulling instruments for exozodiacal disc detection. We have limited our comparison to instruments working at similar wavelengths (ranging from 2 to 8 $\\mu$m), and purposely discarded ground-based instruments working in the N-band such as the KIN and the LBTI. The ultimate performance of these two mid-infrared instruments essentially depends on the spatial and temporal fluctuations of the sky and instrumental thermal backgrounds, which are very difficult to model with a sufficient accuracy for our comparative study. ", "conclusions": "Nulling interferometry is a promising technique to assess the level of circumstellar dust in the habitable zone around nearby main sequence stars. From the ground, instruments like GENIE (VLTI nuller, using two 8-m telescopes) and ALADDIN (Antarctic nuller, using two dedicated 1-m telescopes) could achieve the detection of exozodiacal discs with a density of several tens of zodis. The high Antarctic plateau is a particularly well suited site in that context, so that ALADDIN is expected to achieve the best sensitivity (down to 30 zodis in few hours of integration time). Observing from space provides the solution to go beyond this sensitivity by getting rid of the high thermal background constraining ground-based observations. In this paper, we have investigated the performance of two space-based nulling interferometers which have been intensively studied during the past few years (namely \\peg{} and FKSI). Even though they have been initially designed for the characterisation of hot extrasolar giant planets, \\peg{} and FKSI would be very efficient to probe the inner region of circumstellar discs where terrestrial habitable planets are supposed to be located. Within a few minutes, \\peg{} (resp.\\,FKSI) could detect exozodiacal discs around nearby main sequence stars down to a density level of 5 (resp.\\,1) times our solar zodiacal cloud and thereby outperform any ground-based instrument. FKSI can achieve this sensitivity for most targets of the \\darwin/TPF catalogue while \\peg{} becomes less sensitive for the closest targets with detectable density levels of about 40 times the solar zodiacal cloud. This outstanding and uniform sensitivity of FKSI over the \\darwin/TPF catalogue is a direct consequence of the short baseline length (12.5\\,m) used in combination with an appropriate observing wavelength of about 8\\,$\\mu$m, which is ideal for exozodiacal disc detection. Another advantage of FKSI is to be relatively insensitive to the uncertainty on stellar angular diameters, which is a crucial parameter driving the performance of other nulling interferometers. In terms of sky coverage, we show that these space-based instruments are able to survey about 50\\% of the \\darwin/TPF target stars. The sky coverage reaches 80\\% if they are used in combination with ALADDIN, which provides a complementary sky coverage. Beyond the technical demonstration of nulling interferometry in space, the present study indicates that \\peg{} and FKSI would be ideal instruments to prepare future life-finding space missions such as \\darwin/TPF. \\begin{figure}[t] \\centering \\includegraphics[width=9.2cm]{sky_coverage.eps} \\caption{Sky coverage after 1 year of observation of GENIE (dark frame), ALADDIN (light frame) and \\peg{} (shaded area) shown with the \\darwin/TPF all sky target catalogue. The blue-shaded area shows the sky coverage of a space-based instrument with an ecliptic latitude in the [-30\\degre,30\\degre] range (such as \\peg). The sky coverage of FKSI is similar to that of \\peg{} with an extension of 40\\degre\\,instead of 60\\degre.}\\label{Fig:skycov} \\end{figure}" }, "0808/0808.1716_arXiv.txt": { "abstract": "In the solar convection zone, rotation couples with intensely turbulent convection to drive a strong differential rotation and achieve complex magnetic dynamo action. Our sun must have rotated more rapidly in its past, as is suggested by observations of many rapidly rotating young solar-type stars. Here we explore the effects of more rapid rotation on the global-scale patterns of convection in such stars and the flows of differential rotation and meridional circulation which are self-consistently established. The convection in these systems is richly time dependent and in our most rapidly rotating suns a striking pattern of localized convection emerges. Convection near the equator in these systems is dominated by one or two nests in longitude of locally enhanced convection, with quiescent streaming flow in between at the highest rotation rates. These active nests of convection maintain a strong differential rotation despite their small size. The structure of differential rotation is similar in all of our more rapidly rotating suns, with fast equators and slower poles. We find that the total shear in differential rotation $\\Delta \\Omega$ grows with more rapid rotation while the relative shear $\\Delta \\Omega/ \\Omega_0$ decreases. In contrast, at more rapid rotation the meridional circulations decrease in energy and peak velocities and break into multiple cells of circulation in both radius and latitude. ", "introduction": "Our sun is a magnetic star whose cycles of magnetic activity must arise from organized dynamo action in its interior. This dynamo action is achieved by turbulent plasma motions in the solar convection zone, which spans the outer 29\\% of the sun in radius. Here vigorous convective motions and rotation couple to drive the differential rotation and meridional circulation. These global-scale flows are important ingredients in stellar dynamo theory, providing shear which may build and organize fields on global scales. When our sun was younger, it must have rotated much more rapidly, as is suggested both by the solar wind which continually removes angular momentum from the sun and by many observations of rapidly rotating solar-like stars. In more rapidly rotating suns the coupling between rotation and convection is strong and must continue to drive global scales of flow. Understanding the nature of convection, differential rotation and meridional circulation in more rapidly rotating stars is a crucial step towards understanding stellar dynamos. The manner in which the sun achieves global-scale dynamo action is gradually being sorted out. Helioseismology, which uses acoustic oscillations to probe the radial structure of the star as well as the convective flows beneath the surface, has revealed that the solar differential rotation profile observed at the surface prints throughout the convection zone with two prominent regions of radial shear. The near-surface shear layer occupies the outer 5\\% of the sun, whereas a tachocline of shear at the base of the convection zone separates the strong differential rotation of that zone from the uniform rotation of the deeper radiative interior \\citep[e.g.,][]{Thompson_et_al_2003}. The solar global magnetic dynamo, responsible for the 22-year cycles of activity, is now believed to be seated in that tachocline. In such interface dynamo models, magnetic fields generated in the bulk of the convection zone are pumped into the tachocline where the radial shear builds strong toroidal magnetic fields, with magnetic buoyancy leading to loops that rise upward and erupt through the solar surface \\citep[e.g.,] []{Charbonneau_2005,Miesch_2005}. The differential rotation plays an important role in building and organizing the global-scale fields while the meridional circulations may be important for returning flux to the base of the convection zone and advecting it equatorward, enabling cycles of magnetic activity. By studying the coupling of rotation and convection over a range of conditions, we are likely to learn about the operation of the current solar dynamo and about the nature of dynamos operating in our sun's past and in other solar-like stars. Observations of young solar-like stars indicate that they rotate as much as 50 times faster than the current solar rate. Many of these more rapidly rotating suns possess strong magnetic fields. A correlation between rotation and magnetic activity is observed over a range of stellar types and populations, indicating that more rapidly rotating stars may have stronger stellar dynamos \\citep[e.g.,][]{Noyes_et_al_1984a,Patten&Simon_1996}. Probing the nature of these dynamos and the impact of faster rotation on the internal stellar dynamics requires both accurate observations and detailed dynamical models of the stellar interiors. The faster flows of differential rotation are much easier to detect than the relatively slow motions associated with meridional circulations; observations across the HR diagram indicate that differential rotation is a common feature in many stars. Asteroseismic observations with the Kepler and Corot missions may soon begin to constrain the internal rotation structure. At present only measurements of surface differential rotation are available, as assessed with a variety of techniques including photometric variability \\citep{Donahue_et_al_1996, Walker_et_al_2007}, Doppler imaging \\citep{Donati_et_al_2003} and Fourier transform methods \\citep{Reiners&Schmitt_2003}. Advances in supercomputing have enabled 3-D simulations that are beginning to capture many of the dynamical elements of the solar convection zone. Early global-scale simulations of solar convection by \\citet{Gilman_1975,Gilman_1977,Gilman_1979} under the Boussinesq approximation were extended by the pioneering work of \\citet{Gilman&Glatzmaier_1981}. Such global-scale simulations of solar convection conducted in full spherical shells sought to capture the largest scales of convective flows and began to study how they can establish differential rotation and meridional circulations. However, the range of spatial and temporal scales present in solar convection are vast and thus the computational resources required by the modeling are daunting. Through recent advances in massively parallel computer architectures, solar convection simulations are now beginning to make detailed contact with the observational constraints provided by helioseismology \\citep[e.g.,][]{ Brun&Toomre_2002, Miesch_et_al_2006, Miesch_et_al_2008}. Other efforts have focused on the vigorous turbulence and the dynamo action achieved in the bulk of the solar convection zone \\citep{Brun_et_al_2004}, with recent studies beginning to include the tachocline as a region of penetrative overshoot, shear, and magnetic field amplification \\citep{Browning_et_al_2006}. Facilitated by these computational advances, models of convection and dynamo action within the cores of A-type stars have also begun to be investigated \\citep{Browning_et_al_2004, Brun_et_al_2005, Featherstone_et_al_2007}. To date, most models of stellar differential rotation in stars like our sun that rotate more rapidly have been carried out in 2-D under the simplifying assumptions of mean field theory \\citep[e.g.,][]{Rudiger_et_al_1998, Kuker&Stix_2001, Kuker&Rudiger_2005_A&A, Kuker&Rudiger_2005_AN}. The time is ripe to pursue the question with fully 3-D simulations of global-scale stellar convection. In order to study solar-like stars that rotate faster, our previous work focused on a series of 3-D compressible simulations within a spherical shell using the anelastic spherical harmonic (ASH) code for stars rotating from one to five times the current solar rate \\citep{Brown_et_al_2004}. These preliminary hydrodynamic simulations explored how stellar convection changes with more rapid rotation, including the differential rotation and meridional circulation that is achieved. Comparable studies have been carried out by \\citet{Ballot_et_al_2007} studying younger stars with deeper convection zones. In our simulations we found that remarkable nests of vigorous convection emerge in the equatorial regions. Namely, convective structures at low latitudes about the equator can exhibit strong spatial modulation with longitude, and at high rotation rates the convection is confined to narrow intervals (or nests) in longitude. In the more turbulent simulations presented in this paper, which span a larger range of rotation rates from one to ten times the current solar rate, the phenomena of modulated convection has persisted and the active nests of convection are prominent features at the higher rotation rates. These nests of localized convection persist for long intervals of time and despite their small filling factor maintain a strong differential rotation. The emergence of spatially localized convective states has been observed in other systems, particularly in theoretical studies of doubly-diffusive systems such as thermosolutal convection \\citep[e.g.,][]{Spina_et_al_1998, Batiste_et_al_2006}, in laboratory studies of convection in binary fluids \\citep[e.g.,][]{Surko_et_al_1991}, and in simulations of magnetoconvection where isolated ``convectons'' have been observed \\citep{Blanchflower_1999}. In shells of rapidly rotating fluid, temporally intermittent patches of localized convection emerged in Boussinesq simulations of the geodynamo \\citep{Grote&Busse_2000}. In many of these systems, spatial modulation occurs in the weakly nonlinear regime close to the onset of convection. In contrast, our simulations of solar convection are in a regime of fully developed turbulent convection. We describe briefly in \\S\\ref{sec:ASH} the 3-D anelastic spherical shell model and the parameter space explored by our simulations. In \\S\\ref{sec:convection} we discuss the nature of convection realized in more rapidly rotating stars and the emergence of spatially-localized patterns of convection. In \\S\\S\\ref{sec:global scale flows}-\\ref{sec:energies} we examine the global-scale flows realized in our simulations, including differential rotation and meridional circulation, and their scaling with more rapid rotation. A detailed exploration of the active nests of convection is presented in \\S\\ref{sec:patches}. We reflect in \\S\\ref{sec:conclusion} on the significance of our findings. ", "conclusions": "\\label{sec:conclusion} When stars like our sun are young they rotate much more rapidly than the present sun. In these stars rotation must strongly influence the convective motions and may lead to stronger global-scale dynamo action. We have explored here the effects of more rapid rotation on global-scale convection in simulations of stars like our sun. The mean zonal flows of differential rotation become much stronger with more rapid rotation, scaling as $\\Delta \\Omega \\propto \\Omega_0^{0.3}$ or as $\\Delta \\Omega/\\Omega_\\mathrm{eq} \\propto \\Omega_0^{-0.6}$. In striking contrast, the meridional circulations become much weaker with more rapid rotation, and the energy contained in them drops approximately as $\\Omega_0^{-0.9}$. Accompanying the growing differential rotation is a significant latitudinal temperature contrast, with amplitudes of $100~\\mathrm{K}$ or higher in the most rapidly rotating cases. The maximum temperature contrast near the surface occurs between the hot poles and the cool mid latitudes at about $\\pm 40^\\circ$. If this latitudinal temperature contrast prints through the vigorous convection at the stellar surface, it may appear as an observable latitudinal variation in intensity. The thermal contrast would presumably persist for long periods compared to stellar activity, offering a way to disentangle this intensity signature from that caused by spots of magnetism at the stellar poles. These simulations are entirely hydrodynamic and this provides a major caveat to our findings on the scaling of differential rotation and latitudinal temperature contrast with rotation rate $\\Omega_0$. Prior MHD simulations of stellar convection have demonstrated that in some parameter regimes strong dynamo-generated magnetic fields can react back strongly on the differential rotation, acting to lessen angular velocity contrasts or largely eliminate them \\citep[e.g.,][]{Brun_et_al_2005, Featherstone_et_al_2007,Browning_2008}. It is unclear whether the scaling trends identified here for differential rotation as a function of $\\Omega_0$ will survive in the presence of dynamo action and magnetic fields. Likewise, magnetic fields may lessen the strong temperature contrasts realized here, doing so through their feedback on the convective flows and energy transport. We expect that the weaker meridional circulations may be less affected by magnetic feedbacks, and thus the prediction that meridional circulations lessen in energy and amplitude with more rapid rotation may be of greater significance though harder to confirm observationally. Weaker meridional circulations in more rapidly rotating stars will have a strong impact on many theories of stellar dynamo action, including the Babcock-Leighton flux-transport model favored for solar-type stars. We have begun simulations to explore the dynamo action realized at various rotation rates, and its impact upon the flows described here. Preliminary results appear in \\cite{Brown_et_al_2007c} and more detailed results will be forthcoming shortly. A striking feature of these simulations is the emergence of a pattern of strongly modulated convection in the equatorial regions. These nests of active convection are regions of enhanced convective vigor and transport which propagate at rates distinct from either the mean zonal flows of differential rotation or the individual convective cells. In the most rapidly rotating systems, such as case~G10, convection at the equator is entirely dominated by motions inside the nest with only very weak radial motions present in the regions outside the nest. Though their impact on the convection is most obvious in the rapidly rotating limit, we find some evidence for weak modulation even in our more slowly rotating cases. All of our simulations stop short of the turbulent stellar surface, and it is thus difficult to estimate how these nests of active convection may affect stellar observations in detail. The convective velocities associated with the nests are small compared to the nearly supersonic flows in stellar granulation, and in the sun such global-scale convective structures have evaded direct detection despite intensive searches throughout the near-surface layers by helioseismology. The extremely localized nests found in our most rapidly rotating cases may however influence the thermal stratification and thus convective vigor in the near surface regions, as most of the flux at the equator is transported through a narrow range of longitudes. These nests may act as traveling hot spots with enhanced convection even in the surface layers where the higher emerging flux escapes the system. These spatially localized states of convection may also have some bearing on the active longitudes of magnetic activity observed in the sun, if the enhanced pummeling of the tachocline by the convection within the nest preferentially destabilizes magnetic structures within the tachocline that then rise to the surface. Initial dynamo simulations indicate that nests of convection can coexist with magnetism in portions of parameter space. Thus their strongest signature is likely to emerge in magnetic stars, where magnetic fields threading the bulk of the convection zone may be concentrated in the nests and mimic giant, propagating starspots which survive for very long epochs in time. If the nests lead to active longitudes of enhanced magnetic activity in rapidly rotating stars, we might expect these long-lived magnetic structures to propagate at a rate different from the stellar rotation rate as measured at the surface or from the stellar differential rotation. We recognize that our simulations remain separated by many orders of magnitude from the parameter space of real stellar convection. As such, we must be cautious with our interpretations of the overall dynamics. However, we have found that these nests of convection are a robust feature over a range of parameters and that they are able to persist as entities for as long as we could pursue their modelling. Thus one should be prepared to consider the possibility of their presence also in real stellar convection zones, where they may appear as long-lived propagating features." }, "0808/0808.2838_arXiv.txt": { "abstract": "Here we present a detailed analysis of solar acoustic mode frequencies and their rotational splittings for modes with degree up to 900. They were obtained by applying spherical harmonic decomposition to full-disk solar images observed by the Michelson Doppler Imager onboard the {\\em Solar and Heliospheric Observatory} spacecraft. Global helioseismology analysis of high-degree modes is complicated by the fact that the individual modes cannot be isolated, which has limited so far the use of high-degree data for structure inversion of the near-surface layers ($r > 0.97 R_{\\odot}$). In this work, we took great care to recover the actual mode characteristics using a physically motivated model which included a complete leakage matrix. We included in our analysis the following instrumental characteristics: the correct instantaneous image scale, the radial and non-radial image distortions, the effective position angle of the solar rotation axis and a correction to the Carrington elements. We also present variations of the mode frequencies caused by the solar activity cycle. We have analyzed seven observational periods from 1999 to 2005 and correlated their frequency shift with four different solar indices. The frequency shift scaled by the relative mode inertia is a function of frequency alone and follows a simple power law, where the exponent obtained for the $p$ modes is twice the value obtained for the $f$ modes. The different solar indices present the same result. ", "introduction": "\\label{S-Introduction} The central frequencies of solar acoustic modes, which are obtained using spherical harmonic decomposition, have been successfully used to determine the solar interior structure to as close as 21 Mm to the solar surface ($r < 0.97 R_{\\odot}$) using modes with angular degrees $\\ell \\le 300$ ({\\it e.g.}, \\opencite{Gough96}). The inclusion of high-degree modes ({\\it i.e.}, up to $\\ell = 1000$) has the potential to improve dramatically the inference of the sound speed and the adiabatic exponent ($\\Gamma_1$) in the outermost 2 to 3\\% of the solar radius, allowing to construct localized kernels as close to the solar surface as 1.75 Mm \\cite{Rabello-Soares00}. The effects of the equation of state, through the ionization of hydrogen and helium, are felt most strongly in the outer layers of the Sun, making this shallow region of particular interest. Furthermore, dynamical effects of convection, and the processes that excite and damp the solar oscillations, are predominantly concentrated in this region. Although the spatial resolution of the modern helioseismic instruments allows us to observe oscillation modes up to $\\ell$ = 1000 and higher, only a small fraction of them are currently used ($\\ell \\le 300$). Unfortunately, analysis of high-degree data is complicated by the fact that the individual modes cannot be isolated ({\\it e.g.}, \\opencite{RKS01}). The solar structure is not static, but changes over the solar cycle. It is well known that the mode frequencies change with solar activity. It seems that the {\\color{black} responsible} % mechanism is restricted to the outer layers of the Sun (\\opencite{LW90}), where the high-degree modes are confined. However, at the moment, there is no general agreement in the precise physical mechanism that gives rise to the frequency variation. It is likely a product of the change in the subphotospheric small-scale magnetic field strength with the solar activity cycle ({\\it e.g.}, \\opencite{Goldreich91}). Accordingly to \\inlinecite{DG04}, the frequency shift is easily explained in terms of a variation in the turbulent velocities associated with the magnetic field variation rather than the sole direct effect of the magnetic field itself. \\inlinecite{Li03}, using models of the structure and evolution of the Sun, found that turbulence near the surface of the Sun plays a major role in solar variability, and only a model that includes a magnetically modulated turbulent mechanism can agree with the observed correlation between the frequency shift and the solar cycle. In such a dynamic model, the evolution of the subsurface layers of the Sun {\\color{black} through} % the activity cycle plays an important role. The frequencies of the global modes give the radius and latitude $(r,\\theta)$ part of the structure. While, local helioseismic techniques such as ring-diagram analysis \\cite{Hill88} allow the determination of the three-dimensional structure of the Sun, allowing the study of localized areas in the solar surface, such as those in active regions. Large variations of the mode frequencies observed in and near sunspots in comparison to magnetically quiet regions are well known to be correlated with variations in the average surface magnetic field between the corresponding regions. ({\\it e.g.}, \\opencite{Rajaguru01} and \\opencite{RBB07}). Whether the frequencies are changed directly by the magnetic field or indirectly through an associated change in the solar structure ({\\color{black} such as} % a pressure change) is still a matter of debate. A detailed analysis of the frequency-shift characteristics will hopefully help understand their physical origin. \\inlinecite{Basu04}, using ring-diagram analysis, found that the sound speed is lower in the immediate subsurface layers of an active region than of a magnetically quiet region, while the opposite is true for depths below about 7 Mm. However, \\inlinecite{Basu02}, using global analysis, have found no observable structural changes in the inner layers of the Sun below a depth of 21 Mm associated with the magnetic-activity induced frequency shifts. They were, however, unable to get closer to the solar surface due to the lack of high-degree modes in their mode set. High-degree global analysis is important to complete the picture of the near-surface layers. Besides, the determination of high-degree frequencies using different methods allows us to check the results against each other giving confidence in the results and avoiding systematic errors. We should point out that, although the high-degree modes have short lifetimes (one\\,--\\,ten % hours for $100 \\le \\ell \\le 600$ accordingly to \\opencite{Olga07}) propagating only locally, they are averaged over most of the solar surface using spherical harmonic decomposition (over a relatively long time series) and thus can still be called global analysis. In the traditional global helioseismology data-analysis methodology, a time series of full-disk Doppler solar images is decomposed into spherical harmonic coefficients, characterized by its degree ($\\ell$) and its azimuthal order ($m$). Each coefficient time series is Fourier transformed, and the order of the radial wave function ($n$) gets separated in the frequency domain. However, a spherical harmonic decomposition is not orthonormal over less than the full sphere -- {\\it i.e.}, the solar surface that can be observed from a single view point-- resulting in what is referred as spatial leakage. At low and intermediate degrees, most of these leaks are separated in the frequency domain from the target mode (except for some $m$ leaks) and individual modes can be identified and fitted. However, at high degrees, the spatial leaks lie closer in frequency (due to a smaller mode separation) and, at high frequency, the modes become wider (as the mode lifetimes get smaller), resulting in the overlap of the target mode with the spatial leaks that merges individual peaks into ridges (see Figure~1 in \\opencite{RKS01}). The characteristics of the resulting ridge (central frequency, amplitude, {\\it etc}\\ldots) do not correspond to those of the underlying target mode. This has so far hindered the estimation of unbiased mode parameters at high degrees. To recover the actual mode characteristics, we need a very good estimation of the relative amplitude of the spatial leaks present in a given $(\\ell, m)$ power spectrum, also known as the leakage matrix, which in turn requires a very good knowledge of the instrumental properties \\cite{RKS01}. In our previous papers (\\opencite{RKS01} and \\opencite{KRS04}, hereafter KRS), we described in detail the large influence of the instrumental properties on the amplitude of the leaks and as a consequence in the determination of unbiased high-degree mode parameters. In the following, we will first describe the data used in this analysis and the ridge-to-mode correction applied to them (Sections~\\ref{useddata} and \\ref{ridgemodel}). In Section~\\ref{instreff}, we will discuss the influence on the mode parameters of each of the instrumental properties that were included in the spherical harmonic decomposition of the solar images. We then analyze in Section~\\ref{highdegree} the characteristics of the high-degree mode frequencies and their rotational splittings obtained in this work. Finally, in Section~\\ref{solarcycle}, we analyze the frequency variation induced by the solar cycle. ", "conclusions": "\\label{S-Conclusion} In the determination of unbiased high-degree mode parameters, the instrumental characteristics must be taken into account in the image spatial decomposition or in the leakage matrix calculation itself to obtain a correct estimation of the relative amplitude of the spatial leaks. Among the instrumental characteristics analyzed here, the image scale is the one that affects the parameter determination the most. The image scale is the ratio of the image dimensions observed on the CCD detector and the dimensions in the actual Sun. An error in the image scale introduces an error in the estimated central frequency which increases with the mode frequency. A 0.27\\% error in the image scale would shift the estimated central frequency by as much as 11 $\\mu$Hz at 5 mHz. The radial distortion also has an important effect which is expected since it is similar to an image scale error. An instrumental property not taken into account here (due to a lack of a good estimation) that could have an important effect on the measured parameters is an azimuthally varying PSF. The applied ridge-to-mode correction recovers frequencies at moderate degree that differ from the assumed corrected values by 1 $\\mu$Hz or less depending on the mode frequency. The fitting uncertainty of the recovered frequencies is in the range 0.07\\,--\\,0.18 $\\mu$Hz. The agreement for the $a_1$, $a_2$, and $a_3$ coefficients is very good, except maybe for $f$- and $p_1$-mode $a_1$ coefficients, their mean difference with respect to the assumed correct values is, respectively, three and two normalized by the ridge fitting uncertainties. At high degree, the differences between our frequency determination and theoretical frequencies for the $p$ modes {\\color{black}shows} the same general variation with degree as the results obtained with ring analysis. For $n \\leq 5$, the global and ring analysis agree within 6 $\\mu$Hz. The high-degree $f$-mode frequencies obtained using ring analysis, like previous observations (\\opencite{Duvall98} and references within), are substantially lower than the theoretical frequencies. Surprisingly, the $f$-mode high-$\\ell$ set frequencies agree well with the model frequencies (within 3 $\\mu$Hz) whereas the ring-analysis frequency differences can be as large as 13 $\\mu$Hz for $\\ell > 700$ modes. The implications of the high-$\\ell$ frequencies and splitting coefficients on the solar structure and rotation will be addressed in a future paper. {\\color{black}As noted} % by other authors for low- and moderate-degree modes ({\\it e.g.}, \\opencite{LW90}), the frequency shift induced by the solar cycle scales well with the mode inertia. We extended this analysis to high-degree modes and found out that scaling with the mode inertia normalized by the inertia of a radial mode of the same frequency follows a simple power law (given by Equation~\\ref{eq:plaw}) with one exponent at all frequency ranges, where the $f$ and $p$ modes are fitted independently of one another. The exponents obtained using four different solar indices agree within their fitting uncertainty, where: $\\bar{\\gamma_f} = 1.29 \\pm 0.07$ and {\\color{black}$\\bar{\\gamma_p} = 3.60 \\pm 0.01$}. % The $f$-mode exponent is less than half of the $p$-mode value. The fundamental mode of solar oscillations has essentially the character of surface gravity waves and, contrary to the $p$ modes, it is essentially incompressible and independent of the hydrostatic structure of the Sun. Hence, it is not a surprise that these different types of modes have different exponents. Due to their different properties, it is also very likely that different physical effects are responsible for their frequency variation. Accondingly to \\inlinecite{DG05}, for the $f$ modes, the dominant cause of frequency shift is the variation of the subphotospheric magnetic field. For the $p$ modes, it is the decrease in the radial component of the turbulent velocity in the outer layers during the increase in solar activity, which is accompanied by a decrease in temperature (due to a decrease in the efficiency of convective transport). At low frequency ($\\nu < 2.3$ mHz), the $p$-mode frequency shifts have a different behavior than at high-$\\nu$: a step (with the same exponent $\\gamma_p$) or, as found by other authors, an exponent twice as large as the one at high-$\\nu$. Low-frequency modes have a large probability of being uncorrelated with the solar cycle, which was not taken into account in the case where a large exponent was estimated. {\\color{black} The $f$-mode frequency shifts also have a different behavior around 1.7 mHz: they increase abruptly by an order of magnitude. } Modes with frequency around 3 mHz have the smallest probability that their variation is linearly uncorrelated with the solar index, while modes with $\\nu < 2.5$ mHz or $\\nu > 4.5$ mHz have the largest probability of being uncorrelated. A large probability ($P_u$) of being uncorrelated does not necessarily means that a given mode is not physically correlated with the solar cycle, instead it could be due to uncertainties in the measurements, a low signal-to-noise ratio. The logarithm of $P_u$ is well correlated with the logarithm of the relative uncertainty of the estimated frequency shift $\\delta\\nu^e$, the Pearson correlation coefficient is 0.71 for medium-$\\ell$ modes and 0.54 for high-$\\ell$ modes. If a given mode has a large probability of being linearly uncorrelated, it is expected that the linear fitting of its frequency shifts will have a large uncertainty, hence the high correlation coefficient between $P_u$ and the estimated frequency shift uncertainty. However, it raises the question of what could be the physical process that would make those modes less sensitive to solar activity. For a given $\\ell$, the upper reflection point for lower-frequency modes is deeper in the Sun than for high-frequency modes. If the perturbation layer causing the frequency shift is above the upper turning point of the mode, it would not be affected by the solar cycle. Accordingly to model S of Christensen-Dalsgaard, the depth of the upper turning point increases sharply with decreasing frequency below 2.3 mHz. The upper turning point for a radial mode with $\\nu = 2$ mHz is 0.5 Mm deeper in the Sun than a three-millihertz mode (from Figure~2 in \\opencite{Cha01}). At high-frequency, the observed frequency shift seems to suddendly drop to zero. However, there is no agreement on the exact frequency that this happens, the observed values range from 3.7 to 5 mHz. The frequency-shift falloff is explained by an increase of chromospheric temperature and magnetic field at solar maximum (\\opencite{Jain96} and references within). In the presence of an inclined magnetic field, high-frequency modes tunnel through the temperature minimum and are particularly sensitive to changes in the chromosphere, which are expected to be well correlated with solar activity. \\begin{acks} We are grateful to Tim Larson of Stanford University for discussing with us the results of his improved analysis of MDI medium-$\\ell$ data. The Solar Oscillations Investigation (SOI) involving MDI is supported by NASA grant NNG05GH14G at Stanford University. SOHO is a mission of international cooperation between ESA and NASA. SGK is supported by NASA grant NNG05GD58G. NOAA Mg {\\sc ii} Core-to-wing ratio data are provided by Dr. R. Viereck, NOAA Space Environment Center. The solar radio 10.7 cm daily flux (2800 MHz) have been made by the National Research Council of Canada at the Dominion Radio Astrophysical Observatory, British Columbia. The International Sunspot Number was provided by SIDC, RWC Belgium, World Data Center for the Sunspot Index, Royal Observatory of Belgium. This study includes data from the synoptic program at the 150-Foot Solar Tower of the Mt. Wilson Observatory. The Mt. Wilson 150-Foot Solar Tower is operated by UCLA, with funding from NASA, ONR, and NSF, under agreement with the Mt. Wilson Institute. This work utilizes data obtained by the Global Oscillation Network Group (GONG) program, managed by the National Solar Observatory, which is operated by AURA, Inc. under a cooperative agreement with the National Science Foundation. The data were acquired by instruments operated by the Big Bear Solar Observatory, High Altitude Observatory, Learmonth Solar Observatory, Udaipur Solar Observatory, Instituto de Astrof\\'isica de Canarias, and Cerro Tololo Interamerican Observatory. \\end{acks}" }, "0808/0808.0526_arXiv.txt": { "abstract": "{ Progress in understanding the formation and evolution of planetary nebulae (PN) has been restricted by a paucity of well-determined central star masses. To address this deficiency we aim to (i) significantly increase the number of known eclipsing binary central stars of PN (CSPN), and subsequently (ii) directly obtain their masses and absolute dimensions by combining their light-curve parameters with planned radial velocity data. Using photometric data from the third phase of the Optical Gravitational Lensing Experiment (OGLE) we have searched for periodic variability in a large sample of PN towards the Galactic Bulge using Fourier and phase-dispersion minimisation techniques. Among some dozen periodically variable CSPN found, we report here on three new eclipsing binaries: M 3-16, H 2-29 and M 2-19. We present images, confirmatory spectroscopy and light-curves of the systems. } ", "introduction": "The key parameter for any star is its mass. The least model-dependent method to obtain masses is via Keplerian orbits in binary systems. Eclipsing binaries provide critical information necessary to secure the masses and other fundamental stellar parameters. The mass is obtained from radial velocity (RV) orbits of both stars combined with the orbital inclination from photometric eclipses. Yet directly derived masses have not been fully exploited in the case of central stars of planetary nebulae (CSPN), partly because so few eclipsing binary systems are known. Indirect methods, such as nebula modelling (e.g. Gesicki et al. 2006) or model atmosphere fitting (e.g. Napiwotzki 1999), offer a poor alternative to relatively model-independent techniques such as Keplerian orbits. Obtaining direct masses for a sufficiently large number of CSPN is key to understanding the link between the AGB phase of stellar evolution and the final white dwarf stage. Such a study will also greatly improve our understanding of PN shaping mechanisms (e.g. Zijlstra 2007) and can provide compelling candidates for SN Ia progenitors (Tovmassian et al. 2008). Because most stars of low-intermediate mass experience the PN stage, PN are also major contributors to ISM enrichment and the chemical evolution of galaxies. De Marco et al. (2008, and references therein) summarised the status of photometrically variable CSPN, including 12 close binaries with known periods. Three of these are eclipsing: UU Sge (Pollaco \\& Bell 1993), V477 Lyr (Pollaco \\& Bell 1994), and BE UMa (Ferguson et al. 1999). All three have viable mass determinations using both RV and light-curves, with 0.5--0.7 M$_\\odot$ for the hot 60--120 kK primary and 0.15--0.36 M$_\\odot$ for the irradiated 5--7 kK secondary. Their light-curves display strong irradiation effects, in addition to typical emission lines from the heated hemisphere of the secondary. All these systems are believed to be post-CE (common-envelope) binaries, where the secondary spirals in through the primary's AGB envelope. The remaining stars of De Marco et al. 2008 are non-eclipsing and therefore any mass estimates for them are subject to considerable uncertainty. Either RV or photometric monitoring surveys may be used to \\emph{find} CSPN binaries. However, if the ultimate goal (as in this work) is to get the masses from binaries, then it is easier and more efficient to start with photometric surveys to preselect eclipsing binaries, for which RV orbits combined with photometrically-derived inclinations will secure the masses. Indeed, RV monitoring surveys have proven to be difficult with many candidates showing RV variability without definitive periodicity (e.g. De Marco et al. 2004). This paper introduces a novel use of extant online photometric data from the OGLE microlensing survey. We describe in Section 2 a search for periodic variability in over 300 PN towards the Galactic Bulge. In Section 3 we present the discovery of three new eclipsing binary CSPN. We discuss the nature of these new discoveries in Section 4 and conclude in Section 5. ", "conclusions": "We have doubled the known population of eclipsing binary CSPN after searching OGLE-III photometry in a large PN catalogue towards the Galactic Bulge. The post-CE CSPN show bipolar morphologies consistent with the current hypothesis that close binaries lead to non-spherical nebulae. A following paper will present more new close binary CSPN, discuss the selection effects of the search in detail and provide a new, independent estimate the binary fraction of PN. Planned high-resolution spectroscopy will refine system parameters and derive masses of our new eclipsing binaries." }, "0808/0808.2465_arXiv.txt": { "abstract": "We present Spitzer infrared (IR) photometry and spectroscopy of the lensed Lyman break galaxy (LBG), MS1512-cB58 at $z=2.73$. The large (factor $\\sim 30$) magnification allows for the most detailed infrared study of an $L^*_{UV}(z=3)$ LBG to date. Broadband photometry with IRAC (3-10 $\\mu$m), IRS (16 $\\mu$m), and MIPS (24, 70 \\& 160 $\\mu$m) was obtained as well as IRS spectroscopy spanning 5.5-35 $\\mu$m. A fit of stellar population models to the optical/near-IR/IRAC photometry gives a young age ($\\sim9$ Myr), forrming stars at $\\sim 98$ M$_{\\odot}$ yr$^{-1}$, with a total stellar mass of $\\sim 10^9$ M$_{\\odot}$ formed thus far. The existence of an old stellar population with twice the stellar mass can not be ruled out. IR spectral energy distribution fits to the 24 and 70 $\\mu$m photometry, as well as previously obtained submm/mm, data give an intrinsic IR luminosity $L_{IR} = 1-2 \\times 10^{11}$ L$_{\\odot}$ and a star formation rate, SFR $\\sim20-40$ M$_{\\odot}$ yr$^{-1}$. The UV derived star formation rate (SFR) is $\\sim3-5$ times higher than the SFR determined using $L_{IR}$ or $L_{H\\alpha}$ because the red UV spectral slope is significantly over predicting the level of dust extinction. This suggests that the assumed Calzetti starburst obscuration law may not be valid for young LBGs. We detect strong line emission from Polycyclic Aromatic Hydrocarbons (PAHs) at 6.2, 7.7, and 8.6 $\\mu$m. The line ratios are consistent with ratios observed in both local and high redshift starbursts. Both the PAH and rest-frame 8 $\\mu$m luminosities predict the total $L_{IR}$ based on previously measured relations in starbursts. Finally, we do not detect the 3.3 $\\mu$m PAH feature. This is marginally inconsistent with some PAH emission models, but still consistent with PAH ratios measured in many local star-forming galaxies. ", "introduction": "Much of the global star formation at $z>2$ occurs in ultraviolet-luminous Lyman Break Galaxies \\citep[LBGs,][]{steidel96,reddy05}. For most LBGs, much of what we know about their star-formation (star formation rates, dust reddening) is derived from rest-frame UV properties, where considerable degeneracies exist (ie. dust reddening and starburst age). In most $L^*_{UV}$ LBGs, the majority of the UV photons are absorbed by dust \\citep{adelberger00,reddy08}, which then emits the energy in the infrared. Therefore, an accurate measurement of the total infrared luminosities would give a complete census of the reprocessed UV photons and thus, a better determination of the star formation rates in LBGs. LBGs can be detected in the infrared at 24 $\\mu$m with the {\\it Spitzer Space Telescope}. Observations in the $Spitzer$ 24 $\\mu$m bandpass are the most sensitive towards detecting dust obscured star formation at $z<3$. However, even the deepest $Spitzer$ surveys ($f_{24}>20$ $\\mu$Jy, or $L_{IR}\\geqsim3\\times10^{11}$ L$_{\\odot}$ at $z=3$), can not detect the majority of LBGs \\citep[see IR luminosity function of][]{reddy08}. Between $11$ \\citep{yan05,menendez-delmestre07,sajina07,desai07,pope08}, and even an extremely luminous and dusty LBG \\citep{huang07}. However, there have been no mid-IR spectra obtained of LBGs with typical luminosities ($M_{1500} \\sim -21$, $L_{IR}\\sim10^{11}$ L$_{\\odot}$) and dust extinction ($0.01$. Although cB58 is a ``typical'' LBG in terms of its UV and IR luminosities, the lack of a Balmer break in the SED indicates that it is much younger ($t_{age} \\sim 10 Myr$) than a large majority of LBGs. As seen in previous studies \\citep{shapley05}, stellar population fits to the combined optical/near-IR/IRAC photometry give similar parameters to fits to the optical/near-IR photometry alone, but with smaller error bars. The mid-IR photometry do not significantly constrain the mass of an older stellar population, as the rest-frame near IR flux from the current burst ($M_{burst} \\sim 10^9$ M$_{\\odot}$) can dwarf the flux from an older stellar population with twice the mass. The far-IR photometry is reasonably well fit by local starburst templates and gives an $L_{IR} = 1-2\\times10^{11}$ L$_{\\odot}$. In addition, the PAH luminosities and rest-frame 8 $\\mu$m luminosities agree with the $L_{PAH}$-to-$L_{IR}$ and $L_{8}$-to-$L_{IR}$ relations measured in local starbursts of comparable luminosity and high redshift, higher luminosity starbursts. The inferred SFR from the infrared luminosity is consistent with the SFR derived from the extinction corrected $H\\alpha$ luminosity, but the UV-derived (extinction corrected) SFR is larger by a factor of 3-5. This phenomenon has been noted by \\citet{reddy06}, where LBGs with starburst ages less than 100 Myr consistently have lower $L_{IR}$ than the UV spectral slope and luminosity suggest. With cB58, we can be confident that this discrepancy is {\\it not} due to a poor determination of the infrared luminosity. This suggests that the Calzetti obscuration law for starbursts may not be valid in very young, $<100$ Myr old, high redshift starbursts. The high HI column density and large covering fraction of neutral gas suggests a dust geometry approximated by a uniform foreground sheet. This geometry would result in a steeper reddening law like the LMC or SMC curves and would explain the overstimate of the UV obscuration using the Calzetti reddening law. In addition, because many young ($t_{age}<100$ Myr) LBGs also exhibit spectral properites indicative of large covering fractions of neutral gas (eg. high equivalent width low ionization metal absorption lines), a steeper reddening law may also be appropriate for explaining the relatively red UV spectral slopes observed in these systems as well. It is of course impossible to apply the relations measured in cB58 to the LBG population as a whole. For example, there is significant scatter (beyond measurement errors) in the predicted opacities versus UV spectral slope in the starburst sample used to derive the Calzetti law \\citep{calzetti94}. Therefore, cB58 may simply lie to one end of the natural dispersion already observed in low redshift starbursts. However, the fact that the phenomenon of the UV overpredicting $L_{IR}$ has been observed in a large sample of young LBGs \\citep{reddy06} suggests that cB58 may be typical of these young sytems, and that a systematic bias in UV-derived SFR derivations may exist for the youngest LBGs. SED fits to rest-frame UV and optical photometry of LBGs suggest that at least $\\sim 30$\\% have starburst ages less than 100 Myr \\citep{shapley01,papovich01} and, at least for LBGs with $L_{UV}\\sim L^*$, there is no correlation between age and observed UV luminosity \\citep{shapley01}. Therefore, if the SFR in all of these young LBGs is overestimated by a factor of $\\sim 4-5$, then the global star formation rate density of all LBGs will be overestimated by a factor of $\\sim 2$. $Spitzer$ observations of recently discovered lensed LBGs \\citep{smail07,coppin07,allam07} will allow us to better determine if these discrepancies persist in the rest of the young LBG population." }, "0808/0808.2471_arXiv.txt": { "abstract": "In an earlier paper we quantified the mean merger rate of dark matter haloes as a function of redshift $z$, descendant halo mass $M_0$, and progenitor halo mass ratio $\\xi$ using the Millennium simulation of the $\\Lambda$CDM cosmology. Here we broaden that study and investigate the dependence of the merger rate of haloes on their surrounding environment. A number of local mass overdensity variables, both including and excluding the halo mass itself, are tested as measures of a halo's environment. The simple functional dependence on $z$, $M_0$, and $\\xi$ of the merger rate found in our earlier work is largely preserved in different environments, but we find that the overall amplitude of the merger rate has a strong positive correlation with the environmental densities. For galaxy-mass haloes, we find mergers to occur $\\sim 2.5$ times more frequently in the densest regions than in voids at both $z=0$ and higher redshifts. Higher-mass haloes show similar trends. We present a fitting form for this environmental dependence that is a function of both mass and local density and is valid out to $z=2$. The amplitude of the progenitor (or conditional) mass function shows a similar correlation with local overdensity, suggesting that the extended Press-Schechter model for halo growth needs to be modified to incorporate environmental effects. ", "introduction": "\\label{introduction} In studies of cosmological structure formation, the mass of a dark matter halo is a key variable upon which many properties of galaxies and their host haloes depend. For instance, dark matter haloes of lower mass are expected to form earlier on average than more massive haloes in hierarchical cosmological models such as $\\Lambda$CDM. In semi-analytical modelling of galaxy formation (see \\citealt{BaughReview} for a review), properties such as the formation redshift, halo occupation number, galaxy colour and morphology, and stellar vs AGN feedback processes are all assumed to depend on the mass of the halo (sometimes better characterised by the halo circular velocity). In addition to the halo mass, however, recent work based on numerical simulations has shown that a halo's local environment also affects various aspects of halo formation. For instance, at a {\\it fixed} mass, older haloes are found to cluster more strongly than more recently formed haloes \\citep{Gottlober01, ShethTormen04, Gao2005, Harker06, Wechsler06, JingSutoMo07, WangMoJing07, GaoWhite07, Maulbetsch07}. Other halo properties such as concentration, spin, shape, and substructure mass fraction have also been found to vary with halo environment (e.g., \\citealt{Avila05, Wechsler06, JingSutoMo07, GaoWhite07, Bett07}). In contrast, no such environmental dependence is predicted in the extended Press-Schechter (EPS) and excursion set models \\citep{PS74, BondEPS, LC93} that are widely used for making theoretical predictions of galaxy statistics and for Monte Carlo constructions of merger trees. The lack of environmental correlation arises from the Markovian nature of the random walks in the excursion set model. This limitation is not built into the model per se, but is an assumption stemming from the use of a tophat Fourier-space window function. When a Gaussian window function is used, for instance, \\citet{ZentnerEPS} finds an environmental dependence in the halo formation redshift, but the dependence is {\\it opposite} to that seen in the numerical simulations cited above. Other attempts at incorporating environmental effects into the excursion set model thus far have not been able to reproduce the correlations in simulations (e.g., \\citealt{Sandvik07, DesJacques07}). In this paper, we focus on the environmental dependence of the merger rate of dark matter haloes, a topic that has not been studied in detail. The merger rate is an important quantity for understanding and interpreting observational data on galaxy formation, growth, and feedback processes. While the mergers of galaxies and the mergers of dark matter haloes are not identical processes, the two processes are closely related, and quantifying the latter is the first key step in understanding the former. There have been few theoretical studies of merger rates (e.g., \\citealt{Gottlober01, FM08, Stewart08}) probably because mergers are two- (or more-) body processes, and a large ensemble of descendent haloes {\\it and} their progenitor haloes must be identified from merger trees before the rate can be reliably calculated. In comparison, studies of halo properties such as the mass function, density and velocity profiles, concentration, triaxiality, spin, and substructure distribution require only the particle information from a single simulation output. This paper is an extension of our earlier study \\cite{FM08} (henceforth FM08). There we quantified the global mean merger rates of haloes in the Millennium simulation \\citep{Springel05} over a wide range of descendant halo mass ($10^{12} \\la M_0 \\la 10^{15} M_\\odot$), progenitor mass ratio ($10^{-3} \\la \\xi \\le 1$), and redshift ($0 \\le z \\la 6$). We found that when expressed in units of the mean number of mergers {\\it per halo} per unit redshift, the merger rate has a very simple dependence on $M_0$, $\\xi$, and $z$: the rate depends very weakly on halo mass ($\\propto M_0^{0.08}$) and redshift, and scales as a power law in the progenitor mass ratio ($\\propto \\xi^{-2.01}$) for minor mergers ($\\xi \\la 0.1$), with a mild upturn for major mergers. These simple trends allowed us to propose a universal fitting form for the mean merger rate that is accurate to 10-20\\%. Here we go beyond the global merger rate and use the rich halo statistics in the Millennium database to quantify the merger rate as a function of halo environment, in addition to descendant mass, progenitor mass ratio, and redshift. We also investigate the environmental dependence of the progenitor (or conditional) mass function. This quantity is closely related to the merger rate and is also the most important ingredient in the EPS and excursion set models for constructing Monte Carlo merger trees. Several recent environmental studies have used halo clustering, quantified by the halo bias, as a measure of environment (e.g. \\citealt{Gottlober02, ShethTormen04,Gao2005, Harker06, JingSutoMo07, Wechsler06, GaoWhite07}). While halo bias is a powerful statistical quantity, we choose a simpler and more intuitive local environment measure and use the local mass density centred at each halo. The earlier studies that have used local overdensities as measures of halo environment have used a variety of definitions, e.g., the mean density within a sphere of some radius (ranging from 4 to $10\\, h^{-1}$ Mpc) or within a spherical shell (e.g. between 2 and $5\\, h^{-1}$ Mpc) \\citep{LemsonKauffmann, Harker06, WangMoJing07, Maulbetsch07, Hahn08}. In this paper we compare different definitions of the local overdensity, both including and excluding the mass of the central halo itself. In this paper we also provide an in-depth investigation of the effects of halo fragmentation on the merger rate and its environmental dependence. In FM08, we discussed how fragmentation is a generic feature of all merger trees and compared the {\\it stitching} method with the conventional {\\it snipping} method for handling these events. We will show here that fragmentation occurs more frequently in dense regions than in voids; understanding the effects of fragmentation on merger rates is therefore essential for obtaining robust results in dense environments. There are three general types of approaches to handling fragmentations: do nothing ({\\it snipping}), {\\it stitching} together fragmented haloes, or {\\it splitting} up the common progenitor of the fragmented haloes. We will compare five algorithms for handling fragmentations based on these three approaches and show that except for one algorithm, all the algorithms give similar merger rates to within 20\\%. This paper is organised as follows. In \\S~\\ref{Definitions} we briefly review how haloes and merger trees are constructed from the particle data in the Millennium simulation. Statistics detailing the distribution of halo mass at different redshifts are summarised in Table~\\ref{table:MassBins}. In \\S~\\ref{MeasuringEnvironment} we compare four local density measures and their distributions in relation to halo mass. Three of the measures use the dark matter mass in a sphere centred at a given halo, either including or excluding the mass of the central halo. The fourth measure is motivated by observables such as luminosity-weighted galaxy counts and uses only the masses of the haloes within a sphere. \\S~\\ref{MainResults} contains the main results of this paper, where we quantify how the merger rate is amplified in denser regions and suppressed in voids for redshifts $z=0$ to 2 over three decades of halo mass ($10^{12} - 5\\times 10^{15} M_\\odot$). A simple power-law fitting function for this environmental dependence is proposed, which can be used in combination with the fit for the global rate presented in FM08. We also show that the progenitor (or conditional) mass function has a similar environmental trend as the merger rate. Even though this is expected given that the two quantities are closely related, this result demonstrates directly that the excursion set model is incomplete. In \\S~5, we present statistics of halo fragmentations, compare five algorithms for handling these events, and illustrate the robustness of the results reported in \\S~4. The Appendix provides a discussion of the self-similarity of the merger rate and its environmental dependence in the context of the choice of mass and environment variables used in the fitting formula. ", "conclusions": "\\label{Conclusions} We have used the dark matter haloes and merger trees constructed from the Millennium simulation to quantify the dependence of halo merger rates on halo environment from redshift $z=0$ to 2. A number of local mass density parameters centred at the haloes, both including and excluding the central halo mass itself, are tested as measures of environment. We have found that $\\dsfof$ defined in equation~(\\ref{delta3}) is a robust measure of the surrounding environment outside of a halo's virial radius. It cleanly subtracts out the contributions to the local density from the central halo and thereby breaks the degeneracy between halo mass and environment for high mass haloes (see Figs.~\\ref{fig:DeltaComparison} and \\ref{fig:DeltaDistribution}). We have found strong and positive correlations in both the halo merger rate and the progenitor mass function with environmental densities. Figs.~\\ref{fig:Bn}-\\ref{fig:CMF} present our main results, where haloes in the densest regions are seen to experience 2 to 2.5 times higher merger rates than haloes in the voids. Such a density dependence can be approximated analytically by multiplying our earlier fitting formula FM08 for the global merger rates (eq.~\\ref{eqn:Bnfit}) by an additional $\\delta$-dependent factor given by equation~(\\ref{eqn:FIT}). This factor is a simple power-law in both the environmental density and halo mass, and it is redshift-independent. The mass dependence is quite weak, indicating that haloes with different masses but similar values of $1+\\dsfof$ experience similar merger rate amplifications. This is intriguing in light of the fact, discussed in Section~\\ref{Detangling}, that these haloes actually reside in different environments. The strong correlations of the halo merger rate and progenitor mass function with environment discussed in this paper have important implications for the analytic Press-Schechter \\citep{PS74} and excursion set models \\citep{BondEPS, LC93}. In this popular formalism, halo growth is modelled by the random walk trajectories of dark matter density perturbations smoothed at decreasing scales. Haloes are identified at scales at which these trajectories first cross some critical density threshold, and the Markovian nature of the model allows one to compute the distribution of these first crossings. This distribution is then mapped onto the number-weighted conditional mass function $\\phi(M,z|M_0,z_0)$ discussed in Section~\\ref{CMF} and plays an important role in the Monte Carlo construction of mock merger tree catalogues (see \\citealt{ZFM08} and references therein). It is generally assumed that the conditional mass function is independent of environment as the excursion set model is Markovian. The Markovian nature of the random walks, however, is not a prediction but rather an assumption resulting from the use of the $k$-space tophat window function to smooth the density perturbations. There have been recent attempts to weaken this assumption or to introduce environmental dependence into other parts of the model \\citep{ZentnerEPS, Sandvik07, DesJacques07}, but these modifications thus far have not been able to reproduce the basic statistical correlation between halo clustering and formation time found in simulation studies: older haloes are more clustered \\citep{Gottlober01, ShethTormen04, Gao2005, Harker06, Wechsler06, JingSutoMo07, WangMoJing07, GaoWhite07, Maulbetsch07}. How do our environmental results for the merger rates tie in with these simulation and EPS studies? We have shown that the amplification of halo merger rates in denser regions persists at all redshifts (up to at least $z=2$). If mergers were the dominant channel for halo growth, our results would imply that for haloes of a {\\it fixed} mass today, those in denser regions should have formed {\\it more recently} than those in void regions. Interestingly, this is exactly opposite to the trend reported in many recent studies that have found older (i.e. earlier forming) haloes to be more clustered than younger haloes. As we will discuss in the next paper (Fakhouri \\& Ma 2008c), these two results are in fact not in conflict once the other important channel for halo mass growth -- the ``diffuse'' accretion of non-halo material (either unresolved or stripped) -- is taken into account. We will quantify the environmental dependence of this component and show that, when combined with the merger rate results presented in this paper, we recover the formation redshift dependence reported in prior simulation studies." }, "0808/0808.0704_arXiv.txt": { "abstract": "\\vspace*{5mm} We discuss the interpretation of the annual modulation signal seen in the DAMA experiment in terms of spin-independent elastic WIMP scattering. Taking into account channeling in the crystal as well as the spectral signature of the modulation signal we find that the low-mass WIMP region consistent with DAMA data is confined to WIMP masses close to $m_\\chi \\simeq 12$~GeV, in disagreement with the constraints from CDMS and XENON. We conclude that even if channeling is taken into account this interpretation of the DAMA modulation signal is disfavoured. There are no overlap regions in the parameter space at 90\\%~CL and a consistency test gives the probability of $1.2\\times 10^{-5}$. We study the robustness of this result with respect to variations of the WIMP velocity distribution in our galaxy, by changing various parameters of the distribution function, and by using the results of a realistic $N$-body dark matter simulation. We find that only by making rather extreme assumptions regarding halo properties can we obtain agreement between DAMA and CDMS/XENON. ", "introduction": "The DAMA collaboration has collected an impressive amount of data in their search for the scattering of weakly interacting dark matter particles (WIMPs) off Sodium Iodine. The combined data from DAMA/NaI (7 annual cycles) and DAMA/LIBRA (4 annual cycles) amounts to a total exposure of 0.82~ton~yr~\\cite{Bernabei:2008yi}, in a field where exposure is measured in units of kg~days. DAMA/LIBRA has now provided further evidence for an annual modulation of the event rate in the energy range between 2 and 6~keVee, the claimed statistical confidence of the positive signal being $8.2\\sigma$~\\cite{Bernabei:2008yi}. The phase of the observed modulation (with maximum on day $144\\pm8$) is in striking agreement with the expectation for a modulation in a WIMP scattering signal due to the rotation of the Earth around the Sun, (expected maximum day 152, June 2nd), see e.g.,~\\cite{Jungman:1995df} for a review. An interpretation of this effect in terms of spin-independent interactions of conventional WIMPs with masses $m_\\chi \\gtrsim 50$~GeV is in direct conflict with the constraints from several experiments looking for direct WIMP detection, most notably with the data from CDMS~\\cite{Ahmed:2008eu} and XENON10~\\cite{Angle:2007uj}, which exclude the WIMP cross section consistent with the DAMA modulation for $m_\\chi\\sim 50$~GeV by many orders of magnitude. In light of this, several alternative explanations of the DAMA annual modulation have been proposed, for example spin-dependent interactions~\\cite{Ullio:2000bv, Savage:2004fn}, light WIMPs with $\\lesssim 10$~GeV masses~\\cite{Bottino:2003iu, Bottino:2003cz, Gondolo:2005hh}, keV scale axion-like dark matter~\\cite{Bernabei:2005ca} (see however, \\cite{Pospelov:2008jk, Gondolo:2008dd}), dark matter interacting only with electrons~\\cite{Bernabei:2007gr}, inelastic WIMP scattering~\\cite{TuckerSmith:2001hy, Chang:2008gd} and mirror dark matter~\\cite{Foot:2008nw}. In this work we reconsider the possibility of spin-independent elastic scattering of light WIMPs with $\\lesssim 10$~GeV masses~\\cite{Bottino:2003iu, Bottino:2003cz, Gondolo:2005hh}, see~\\cite{Bottino:2007qg, Bottino:2008mf, Petriello:2008jj, Feng:2008dz} for recent studies. The original idea is that light dark matter scattering on the relatively light Sodium nuclei in DAMA could deposit enough energy in the detector to give a signal, whereas the scattering of light halo particles off heavier nuclei, such as for example Ge in CDMS or Xe in XENON would lead to energy depositions below the threshold of those detectors. Recently the importance of the so-called channeling effect~\\cite{Drobyshevski:2007zj} in the crystal structure of the experiment has also been emphasized~\\cite{Petriello:2008jj, Bottino:2008mf}. Specific models for WIMPs with $m_\\chi \\sim 10$~GeV have been studied for example in~\\cite{Bottino:2002ry, Bottino:2007qg, Barger:2005hb, Gunion:2005rw}. Here we do not discuss theoretical implications but focus on the phenomenology of direct detection experiments in a model-independent way by assuming that such light WIMPs can provide the correct relic abundance while any direct collider constraints can be evaded. In this region of WIMP masses several experiments~\\cite{Altmann:2001ax, Akerib:2003px, Lin:2007ka, Aalseth:2008rx} exclude WIMP--nucleon scattering cross sections in the range $\\sigma_p \\gtrsim 10^{-40}$~cm$^2$. As we will see in the next pages, once we have included channeling as well as the spectral shape of the DAMA modulation signal, the allowed region of our interest is obtained at much small cross sections, around $\\sigma_p \\sim 10^{-41}$~cm$^2$ and $m_\\chi \\sim 10$~GeV. In this region the most relevant constraints come from XENON~\\cite{Angle:2007uj}, the 2008 Germanium data from CDMS~\\cite{Ahmed:2008eu}, and the 2005 CDMS data on Silicon~\\cite{Akerib:2005kh}. Indeed, as we will discuss, the spectral shape of the DAMA annual modulation restricts $m_\\chi$ and $\\sigma_p$ to a region excluded by these experiments. In our study we elaborate on this result and discuss how robust it is with respect to different assumptions about the dark matter halo of our galaxy. The impact of non-standard halo properties on dark matter direct detection experiments has been discussed by many authors, see for example~\\cite{Belli:2002yt, Fornengo:2003fm, Green:2002ht, Vergados:2007nc}. At a qualitative level, one would expect that smaller velocity dispersions or truncated velocity distributions would seem to favour the dark matter interpretation of the DAMA signal, as they could lead to more events above the low energy threshold of DAMA but below that of other experiments. Furthermore, anisotropies in the velocity dispersion could amplify annual modulation signals. The outline of our work is as follows. In Sec.~\\ref{sec:analysis} we briefly summarise the phenomenology of elastic WIMP scattering in direct detection experiments and give some technical details on our analysis of DAMA, CDMS and XENON data. The results for a standard dark matter halo are presented in Sec.~\\ref{sec:std-halo}. In Sec.~\\ref{sec:nonstd-halo} we consider deviations from the standard assumptions made about the WIMP velocity distribution: we use results from the Via Lactea $N$-body dark matter simulation~\\cite{Diemand:2006ik}, we vary several parameters of the Maxwellian distribution and consider asymmetric velocity profiles. Sec.~\\ref{sec:conclusions} contains our conclusions. In Appendix~\\ref{sec:dama-mod} we comment on the DAMA fit using the annual modulation energy spectrum, and in Appendix~\\ref{app:comparison} we briefly compare our results to the ones from other authors. ", "conclusions": "\\label{sec:conclusions} Prompted by recent results from DAMA/LIBRA which establish the annual modulation of their event rate at the 8.2$\\sigma$ level, we have studied the interpretation of this signal in terms of spin-independent elastic WIMP scattering. We have shown that the energy spectrum of the modulation signal strongly restricts the region of WIMP masses below 10~GeV, confining WIMP masses consistent with the DAMA data close to $m_\\chi \\simeq 12$~GeV. This region is excluded by the limits from CDMS and XENON, and therefore we conclude that even if channeling is taken into account this interpretation of the DAMA modulation signal is disfavoured. Applying a stringent test to evaluate the consistency of DAMA with null-result experiments we find consistency only with a formal probability of $10^{-5}$. We have studied how robust this result is with respect to variations of the WIMP velocity distribution in our galaxy by changing various parameters of the distribution function. We find that decreasing the dispersion of the distribution can somewhat reduce the tension in the fit. Adopting in addition an asymmetric WIMP velocity profile with a larger dispersion in the radial direction than tangential improves the fit considerably. We conclude that in principle it is possible to reconsile DAMA in the considered framework, at the price of rather exotic properties of the DM halo. The question remains whether such halo properties can be realistic at all. We have checked that a WIMP velocity distribution based on the Via Lactea $N$-body dark matter simulation does not improve the fit considerably with respect to the standard Maxwellian halo model. Finally we mention that the negative conclusion on the compatibility of DAMA with CDMS and XENON relies crucially on the energy threshold of the latter two. In particular, a shift in the nuclear recoil energy scale in these experiments may change the conclusion. Indeed, the new measurements of the $\\mathcal{L}_\\mathrm{eff}$ parameter in XENON~\\cite{Sorensen:2008ec} (which has not been implemented in the first arXiv version of this work) made the disagreement between DAMA and XENON somewhat less sever." }, "0808/0808.3707_arXiv.txt": { "abstract": "We run adiabatic $N$-body/hydrodynamical simulations of isolated self-gravitating gas clouds to test whether conformal gravity, an alternative theory to General Relativity, is able to explain the properties of X-ray galaxy clusters without resorting to dark matter. We show that the gas clouds rapidly reach equilibrium with a density profile which is well fit by a $\\beta$-model whose normalization and slope are in approximate agreement with observations. However, conformal gravity fails to yield the observed thermal properties of the gas cloud: (i) the mean temperature is at least an order of magnitude larger than observed; (ii) the temperature profiles increase with the square of the distance from the cluster center, in clear disagreement with real X-ray clusters. These results depend on a gravitational potential whose parameters reproduce the velocity rotation curves of spiral galaxies. However, this parametrization stands on an arbitrarily chosen conformal factor. It remains to be seen whether a different conformal factor, specified by a spontaneous breaking of the conformal symmetry, can reconcile this theory with observations. ", "introduction": "\\label{sec:intro} On the scale of individual galaxies and larger scales, General Relativity requires large amounts of dark matter to describe the dynamics of cosmic structure. Moreover, the late-time acceleration of the Hubble expansion implies the existence of a cosmological constant, a special case of a dark energy fluid which is suggested by more sophisticated models (see \\citealt{copeland06}, for a review). In principle, we can avoid the dark matter and dark energy solutions to the puzzles posed by the astrophysical data by adopting an alternative theory of gravity, which reduces to General Relativity in the appropriate limit. Independently of the dark matter and dark energy problems, a modification of General Relativity is also highly desirable if we ultimately wish to unify gravity with the other fundamental interactions. Possible modifications of the Einstein-Hilbert action proposed in the literature are, among others, (i) the introduction of additional scalar and/or vector fields (e.g., \\citealt{fujii03, beken06}); (ii) the assumption of arbitrary functions $f(R)$ of the Ricci scalar $R$ (e.g., \\citealt{capozz-franc07, nojiri08}); (iii) the introduction of additional dimensions to the four dimensions of the General Relativity spacetime manifold (e.g., \\citealt{maart04}). A different approach was suggested by \\citet{mannh90} who revived Weyl's theory \\citep{weyl18, weyl19, weyl20} as a possible candidate to solve the dark matter and dark energy problems. When the geometry is kept Riemannian, with a null covariant derivative of the metric tensor, we can obtain a milder version of Weyl's gravity, known as conformal gravity. In this theory, we impose a local conformal invariance on the gravitational field action in the curved four-dimensional spacetime. The Einstein-Hilbert Lagrangian density for the gravitational field is chosen on the requirement that the theory of gravity is a second-order derivative theory. In conformal gravity, the Lagrangian density is chosen on the principle of local conformal symmetry which is uniquely satisfied by the action $I_W = -\\alpha\\int d^4x\\sqrt{-g} C_{\\mu\\nu\\kappa\\lambda}C^{\\mu\\nu\\kappa\\lambda}$, where $C_{\\mu\\nu\\kappa\\lambda}$ is the Weyl tensor, $\\alpha$ is a coupling constant, and $g$ is the determinant of the metric tensor $g_{\\mu\\nu}$. Conformal symmetry is garanteed by the invariance of the Weyl tensor to local conformal transformations $g_{\\mu\\nu}(x) \\to \\Omega^2(x)g_{\\mu\\nu}(x)$, where $\\Omega^2(x)$ is an arbitrary conformal factor that can be specified by a spontaneous breaking of the conformal invariance \\citep{edery06, mannh08b}. The theory of gravity implied by the action $I_W$ is a fourth-order derivative theory. The vacuum exterior solution for a static and spherically symmetric spacetime contains the Schwarzschild solution \\citep{MK89}. The weak-field limit is consistent with the Solar System observational data \\citep{mannh07}, unlike claimed in previous investigations \\citep{barabash99, flanagan06, barabash07}. \\cite{mannh93, mannh97} shows that conformal gravity can reproduce the rotation curves of disc galaxies without dark matter. Moreover, conformal gravity can explain the current accelerated expansion of the universe without resorting to a fine-tuned cosmological constant or to the existence of dark energy \\citep{mannh01, mannh03, varieschi08}. Unlike the standard theory, where the universe starts accelerating at redshift $z\\la 1$, conformal gravity predicts that the universe is accelerating at all times. Therefore, observational data probing the Hubble plot at very high redshift (e.g., \\citealt{navia08, wei08}) can be a decisive test. More recently, conformal theory has been proposed as a valid candidate for building a theory of quantum gravity \\citep{mannh08}; in fact, theories based on fourth-order derivative equations of motion have had the long-lasting problem of suffering from the presence of ghosts. \\citet{bender08} have recently shown that this erroneous belief is the result of considering the canonical conjugates ${\\bf p}$ of the generic dynamical variables ${\\bf q}$, when the ${\\bf q}$'s are real, to be Hermitian operators; however, this assumption is incorrect, and when the non-Hermiticity property of the Hamiltonian of these higher-order quantum field theories is correctly taken into account, the states with negative norm disappear. From an astrophysical perspective, however, the form of conformal gravity that has been proposed in the literature currently has two main shortcomings: the abundance of light elements, and the gravitational lensing phenomenology. Conformal gravity nicely avoids the requirement of an initial Big Bang singularity, but it still predicts an early universe sufficiently dense and hot to ignite the light element nucleosynthesis. However, conformal gravity predicts a too slow initial expansion rate. This rate favours the destruction of most of the deuterium produced in the early universe \\citep{knox93, elizondo94} and poses conformal gravity in serious difficulties compared to the standard Big Bang nucleosynthesis. Conformal gravity necessarily requires astrophysical processes for the production of the deuterium currently observed, for example neutron radiative capture on protons in the atmospheres of active stars \\citep{mullan99} or gamma-ray bursts \\citep{inoue03}. However, the processes investigated to date do not seem to be efficient enough. For example, significant production of deuterium in the Galaxy seems to be ruled out \\citep{proda03}. The second open issue is the conformal gravity prediction of gravitational lensing. Early investigations of gravitational lensing in conformal gravity \\citep{walker94, edery98, edery99p, edery99} show that, in the weak-field limit, the deflection angle due to a point mass $M$ is $\\Delta\\alpha=4GM/c^2r - \\gamma r$, where $r$ is the radius of the photon closest approach; $\\Delta\\alpha$ contains the additional term $\\gamma r$ when compared to the General Relativity result. The constant $\\gamma$ has to be positive to fit the galaxy rotation curves, thus implying a repulsive effect in gravitational lensing. It was later realized \\citep{edery01} that the geodesics of photons are independent of the conformal factor $\\Omega^2(x)$ and one can choose an appropriate conformal factor and a radial coordinate transformation to yield attractive geodesics for both massive and massless particles. However, in the strong-field limit, the light deflection might still be divergent or even impossible \\citep{pireaux04a, pireaux04b}. Until these open questions are completely settled, conformal gravity cannot yet be considered ruled out by observations. Moreover, conformal invariance plays a crucial role in elementary particle physics and a viable theory of gravity that includes this property can at least be suggestive of a relevant route towards the unification of the fundamental interactions. From the astrophysical point of view, conformal gravity can be further tested by investigating the formation of cosmic structure. To date, nobody has yet explored how the large-scale structure forms in conformal gravity. If structures form by gravitational instability, as in the standard theory, the development of a cosmological perturbation theory, which is still lacking, becomes inevitable. This theory would enable the comparison of conformal gravity with the spectrum of the Cosmic Microwave Background anisotropies and would provide initial conditions for the simulation of the structure evolution into the non-linear regime. Before building up such a theory, however, it is useful to check whether conformal gravity is able to reproduce the equilibrium configuration of cosmic structures, other than galaxies, without dark matter. \\cite{horne06} has already shown that, if we interpret the observational data of the intracluster medium of X-ray clusters by assuming hydrostatic equilibrium, conformal gravity requires a factor of ten less baryonic mass than inferred from the X-ray surface brightness measures. Here, we extend this analysis by performing hydrodynamical simulations of self-gravitating gas clouds. We modify a standard Tree+SPH code to perform numerical simulations of self-gravitating systems in conformal gravity. The numerical tool we create is extremely relevant if we eventually wish to investigate the formation of the large-scale structure, because we will massively need to resort to numerical integrations when the evolution of the density perturbations reaches the non-linear regime. In Section \\ref{sec:CG}, we review the basic steps leading to the gravitational potential of a static point source in conformal gravity. In Section \\ref{sec:sphere}, inspired by the analysis of \\citet{horne06}, we compute the gravitational potential energy of a spherical system, and in Section \\ref{sec:xray} we use this result to compute the expected mean temperature and the temperature profile of the intracluster medium. In Section \\ref{sec:numerics}, we derive the same results with hydrodynamical simulations of self-gravitating gas clouds. ", "conclusions": "Conformal gravity can explain the rotation curves of disk galaxies and the current accelerated expansion of the universe without resorting to dark matter and dark energy. We have modified a Tree+SPH code to run hydrodynamical simulations of isolated X-ray galaxy clusters to show that conformal gravity does not share the same success on the scales of clusters. These simulations confirm our simple analytic estimates that show that gas clouds with mass $\\sim 10^{12}-10^{13}$~M$_\\odot$, which are typical values of the total mass of the hot gas present in real clusters, remain confined with an equilibrium mean temperature $\\sim 10-100$~keV, ten times larger than the observed temperatures; more dramatically, because of the presence of a linear term in the gravitational potential, at large clustrocentric radius $r$, the gas temperature increases with $r^2$, rather than decreasing as in real systems. Our analysis totally neglects radiative cooling and gas heating from astrophysical sources, as supernovae or active galactic nuclei. The interplay between these processes can in principle provide a way to reconcile conformal gravity with observations. It is however unclear if and how much these processes should be fine-tuned to provide X-ray clusters in agreement with observations. In addition to this topic, we can see two more open issues whose solution might also reconcile conformal gravity with observations: \\begin{enumerate} \\item In conformal gravity, all the matter in the universe is expected to affect the local dynamics. The net effect is to contribute a constant inward acceleration $-GM_0/R_0$ in addition to the gravitational acceleration generated by local sources. Because we included this constant acceleration in our simulations, we could neglect the rest of the universe and impose vacuum boundary conditions. It might be possible that assimilating the gravitational influence of the nearby matter surrounding the X-ray cluster in the constant ``universe'' acceleration $-GM_0/R_0$ is inappropriate: in fact, nearby external matter might decrease the gravitational attraction of the interior matter and hopefully reduce the thermal energy of the gas. To appropriately investigate this effect, we should simulate the dynamics of large-scale structure within a full cosmological context. However, this task is not trivial just because the gravitational field is highly non-local. This investigation would also benefit from the implementation into the numerical simulation of the yet unavailable theory of structure formation. \\item The gravitationl potential we implemented in our simulations derives from a metric where the conformal invariance is broken by an arbitrary choice of the conformal factor $\\Omega^2(x)$. It is unclear whether this choice provides a coordinate system whose physics describes the real world or it is an artifact of the reference frame. It also remains to be seen whether a spontaneous breaking of the conformal invariance, in theories where matter and gravity are conformally coupled \\citep{edery06}, can provide a metric, and thus a gravitational potential, where the observed thermal properties of the intracluster medium can be reproduced. \\end{enumerate} In Section \\ref{sec:intro} we mentioned that the nucleosynthesis of light elements and the phenomenology of gravitational lensing are two open issues that need to be solved before accepting conformal gravity as a viable alternative theory of gravity and cosmology. Here, we have shown that the thermodynamics of X-ray clusters poses a third challenge to this theory." }, "0808/0808.3641_arXiv.txt": { "abstract": "{ We discuss the possibility that high-frequency QPOs in neutron-star binary systems may result from forced resonant oscillations of matter in the innermost parts of the accretion disc, excited by gravitational perturbations coming from asymmetries of the neutron star or from the companion star. We find that neutron-star asymmetries could, in principle, be effective for inducing both radial and vertical oscillations of relevant amplitude while the binary companion might possibly produce significant radial oscillations but not vertical ones. Misaligned neutron-star quadrupole moments of a size advocated elsewhere for explaining limiting neutron star periods could be large enough also for the present purpose.} ", "introduction": "% Quasi-periodic oscillations of \\mbox{X-ray} brightness (QPOs) have been observed in a number of accreting binary systems containing compact objects, both with neutron stars \\citep[see][ for a review]{Kli:2000:ARASTRA:,Bar-Oli-Mil:2005:MONNR:} and with black holes \\citep{Rem-McCli:2006:ARASTRA:}. They can have low frequencies ($\\mathrm{Hz}$) or high frequencies ($\\mathrm{kHz}$). The observed $\\mathrm{kHz}$ frequencies are comparable with the Keplerian and epicyclic frequencies in the inner parts of the accretion disc \\citep{Tor:2005:ASTRN:} and the question arises of whether they might be associated with forced resonant oscillations of the inner disc material. In order to initiate these, some perturbation mechanism would be required. Here, we focus on neutron-star systems and investigate the possibility that gravitational perturbations caused either by the binary companion or by asymmetries of the neutron star might provide this mechanism. (The case of the binary companion could be relevant also for black hole systems.) We note that perturbations coming from the neutron star can only be relevant for the innermost parts of the disc (because of the rapid fall-off of the force with distance) and hence are mainly associated with the picture for $\\mathrm{kHz}$ QPO frequencies rather than with that for the lower frequencies. It is necessary that the mechanism should be a resonant one since otherwise the response produced would certainly be much too small to be of interest. The two types of perturbation (from the neutron star and from the binary companion) clearly induce different behaviours: the frequency of the varying force arising from the influence on the disc of the binary companion is essentially equal to the disc rotation-frequency, whereas the main frequency of the force caused by asymmetries of the neutron star is equal to the difference between the rotation frequencies of the disc and the neutron star \\citep{Petri:2006:ApSS:}. Resonance occurs at those points where one of the intrinsic oscillation frequencies of the disc matter coincides with the forcing frequency. Following the discussion by \\citet{Lan-Lif:1976:Mech:} for test particle motion (which would need to be modified for fluid elements), the growth in amplitude of the oscillations of the particle within the linear regime of forced resonance is given by \\begin{equation} a(t) = \\frac{f_{\\mathrm{p}}}{2m_0\\omega}\\,t\\ , \\label{resonance} \\end{equation} where $f_{\\mathrm{p}}$ is the amplitude of the variations of the force, $\\omega$ is the frequency and $m_0$ is the mass of the particle. This linear regime ends when the oscillation amplitude $a(t)$ becomes large enough so that non-linear phenomena and/or dissipative processes become relevant. Note that $a(t)$ grows linearly with time in this regime and so can become quite large even when the variations in the perturbing force are small. ", "conclusions": "\\label{compar}% Here, we discuss whether the effect of the neutron-star asymmetries or of the binary companion could be large enough to account for excitation of the QPO phenomenon. In our simplified picture (c.f. Eq.\\,(\\ref{resonance})), the excitation time for the amplitude $a$ of resonant oscillations to grow to a particular value is given by \\begin{equation} t_{\\mathrm{ex}} = \\left(\\frac{a}{R}\\right)\\, \\left(\\frac{\\alpha}{\\pi}\\right)\\, \\left(\\frac{f_{\\mathrm{p}}}{f_0}\\right)^{\\!-1}\\, \\tau_{\\sssm{K}}\\,, \\end{equation} where $f_p$ is the amplitude of the perturbing force acting on unit mass, $f_0 = GM_{\\sssm{A}}/R^{2}$ is the main gravitational force from the central object, $\\tau_{\\sssm{K}} = 2\\pi/\\Omega_{\\sssm{K}}$ is the period of circular Keplerian motion at the location being considered and the epicyclic frequency of the perturbation being excited is $\\omega = \\alpha\\Omega_{\\sssm{K}}$. The dimensionless radial and vertical ``epicyclic functions'' satisfy $\\alpha \\le 1$ everywhere. Note that since $\\tau_{\\sssm{K}} \\sim 10^{-3}\\,\\mathrm{s}$, the amplitude amplification in 1 second is $\\sim\\! 10^{3}$. The ratio $a/R$ needs to grow to $> 10^{-3}$ in order to potentially explain the QPO behaviour and it would need to do that within $\\sim\\!10^{3}\\,\\mathrm{s}$ to account for the QPO phenomena seen in atoll sources. In the case of a binary companion, for the radial perturbing force one has \\begin{eqnarray} \\frac{f_{\\mathrm{p}}}{f_0} & \\sim & \\left(\\frac{M_{\\sssm{B}}}{M_{\\sssm{A}}}\\right) \\left(\\frac{R}{d}\\right)^{2} \\nonumber \\\\ & \\sim & 10^{-9} \\left( \\frac{M_{\\sssm{B}}}{0.1\\,M_{\\sssm{A}}} \\right) \\left( \\frac{R}{10^{6}\\,\\mathrm{cm}}\\frac{10^{10}\\,\\mathrm{cm}}{d} \\right)^{2}\\, , \\end{eqnarray} where we have normalised to typical parameter values. This could produce $a/R \\sim 10^{-3}$ at times $t_{\\mathrm{ex}}\\lesssim 10^{3}\\,\\mathrm{s}$, but there is a problem for it producing resonances in the innermost parts of the disc because of having $\\omega_{\\sssm{B}} \\approx\\Omega_{\\sssm{K}}$. The corresponding vertical force is smaller by at least four orders of magnitude and is clearly irrelevant. (Note that the vertical force is a tidal force whereas the radial one is a direct gravitational attraction; also, $\\theta_{\\sssm{B}}$ is probably rather close to $\\pi/2$.) For the neutron star asymmetries, we focus on the case of the misaligned quadrupole moments where $f_{\\mathrm{p}}/f_0$ can be $\\sim 10^{-8}$ in the inner parts of the disc for both radial and vertical oscillations (taking $m_{\\mathrm{quad}} = 10^{-8}\\,M_{\\sssm{A}}$). For this, $a/R$ could reach $10^{-3}$ in $\\lesssim 10^2 \\,\\mathrm{s}$, which encourages further investigation of this scenario. We conclude that at least one of the types of gravitational perturbation considered in this paper might provide a plausible mechanism for inducing kHz QPO behaviour although many details remain to be worked out (in particular concerning the response of the fluid medium and the production of the luminosity variations). The influence of the binary companion could possibly be effective in providing the perturbations but the more likely possibility is that they might be produced by the neutron-star asymmetries. It is striking that the same magnitude for the misaligned quadrupole moment as advocated elsewhere for explaining limiting neutron star periods, seems also to give a plausible mechanism for inducing QPO behaviour. \\vspace{\\baselineskip} \\hrule \\vspace{\\baselineskip}" }, "0808/0808.0254_arXiv.txt": { "abstract": "The unexpected high bump in the UV part of the spectrum found in nearby giant elliptical galaxies, a.k.a. the UV upturn, has been a subject of debate. A remarkable progress has been made lately from the observational side, mainly involving space telescopes. The GALEX UV telescope has been obtaining thousands of giant ellipticals in the nearby universe, while HST is resolving local galaxies into stars and star clusters. An important clue has also been found regarding the origin of hot HB stars, and perhaps of sdB stars. That is, extreme amounts of helium are suspected to be the origin of the extended HB and even to the UV upturn phenomenon. A flurry of studies are pursuing the physics behind it. All this makes me optimistic that the origin of the UV upturn will be revealed in the next few years. I review some of the most notable progress and remaining issues. ", "introduction": "A review on the UV upturn phenomenon may usually start with a following or similar definition: ``a bump in the UV spectrum between the Lyman limit and 2500\\AA\\, is found {\\em virtually in all bright spheroidal galaxies}'' (e.g., Yi \\& Yoon 2004). This seems no longer true! While earlier studies based on a small sample of nearby galaxies led us to think so, a much greater sample from the recent GALEX database appears to disprove it. Only a small fraction of elliptical galaxies show a strong UV upturn and it is generally limited to the brightest cluster galaxies (Yi et al. 2005). This review is about the recent development on this seemingly-old topic. I recycle some of the contents in my earlier review given in the first Hot Subdwarf and Related Objects workshop held in Keele, UK (Yi \\& Yoon 2004). For a more traditional review, readers are referred to the articles of Greggio \\& Renzini (1999) and O'Connell (1999). ", "conclusions": "The UV satellite GALEX is obtaining a valuable UV spectral evolution data for numerous bright cluster galaxies. The apparent trend in redshift vs $FUV-V$ colour seems consistent with the prediction from the single stellar population models. This is comforting while observers feel obliged to build up their database much more substantially in order to make it statistically robust. Two new issues are notable. Firstly, binary population synthesis community feels odd to find that the simplistic single-star population models are found to be good enough. The in-principle more advanced binary population models are obliged to reproduce the observed CMDs of simple populations (globular clusters) before attempting to model galaxies. For example, I am very eager to see their models reproduce the ordinary HB first, before explaining the EHB. Secondly, the enhanced helium hypothesis based on the globular clusters in Milky Way and M87 is a very exciting possibility. The deduced value of the helium abundance seems unphysical to be a global property for the galaxy but may be possible for small systems that are vulnerable to a chemical fluctuation in the proto-galaxy cloud. While a more detailed investigation is called for it may be difficult to be influential to the entire stellar population of a galaxy. For instance, adding all spectral energy distributions of the Milky Way globular clusters would not yield anything close to the spectrum of a UV upturn galaxy. Of course, a metallicity difference may act as an added complication. The secret will be revealed through time and hard work, perhaps very soon." }, "0808/0808.0581_arXiv.txt": { "abstract": "A {\\itshape Chandra} study of pulsar wind nebula around the young energetic pulsar PSR B1509$-$58 is presented. The high resolution X-ray image with total exposure time of 190 ks reveals a ring like feature 10'' apart from the pulsar. This feature is analogous to the inner ring seen in the Crab nebula and thus may correspond to a wind termination shock. The shock radius enables us to constrain the wind magnetization, $\\sigma \\geq 0.01$. The obtained $\\sigma$ is one order of magnitude larger than that of the Crab nebula. In the pulsar vicinity, the southern jet appears to extend beyond the wind termination shock, in contrast to the narrow jet of the Crab. The revealed morphology of the broad jet is coincident with the recently proposed theoretical model in which a magnetic hoop stress diverts and squeezes the post-shock equatorial flow towards the poloidal direction generating a jet. ", "introduction": "Recent X-ray observations revealed that part of pulsar wind nebulae (PWNe) have common structures of ``Torus'' and ``Jet'', as typified by the crab pulsar \\citep{2002ApJ...577L..49H,2000ApJ...536L..81W}, the Vela pulsar \\citep{2003ApJ...591.1157P,2001ApJ...556..380H}, and PSR B1509$-$58 \\citep{2002ApJ...569..878G}. However the physical mechanisms which form the ubiquitous features, especially the jets, are still unclear, because the origin of the jet in the Crab is too compact to discriminate its structure even with the current X-ray observatories. The theoretical approaches toward the axisymmetric PWNe have been aimed to figure out the Crab nebula. The torus in the Crab had been successfully explained by one-dimensional MHD simulations \\citep{1984ApJ...283..694K,1984ApJ...283..710K}. A young rotation powered pulsar is believed to lose its rotating energy via a magnetized particle flow from a pulsar, the so-called pulsar wind. Since the pulsar wind is flowing almost at the speed of light, it is stalled by a termination shock just around the pulsar. When the wind passes through the termination shock, the kinetic energy of the wind is converted into internal energy. Essentially, the post-shock flow is decelerated yielding to the conservation law of particle flux $nvr^2 = \\mathrm{const}$, where $n$ is the number density (that is almost constant), $v$ is the flow velocity and $r$ is the radius from the pulsar. At the same time, the frozen-in condition, $rvB = \\mathrm{const}$, induces the compression of the toroidal magnetic field $B$, in which the internal energy is efficiently converted into synchrotron radiation forming an X-ray torus. In practice the magnetic pressure prevents the deceleration of the post-shock flow, thus the magnetization of the pulsar wind is closely related to the geometric structure of the PWN. In this way, several methods to calculate the magnetization parameter $\\sigma$ have been proposed \\citep{1974MNRAS.167....1R,1984ApJ...283..694K,1994ApJ...435..230G}. Moreover, recent two-dimensional numerical studies suggested that the magnetization seems to control the formation of the jet \\citep{2002MNRAS.329L..34L,2004A&A...421.1063D,2006cosp...36..191B}. Therefore it is essential to evaluate the magnetization parameters to understand the formation process of PWNe. \\begin{figure*} \\begin{center} \\FigureFile(170 mm,120mm){overview_closeup.eps} \\end{center} \\caption{(Left)--- Overview image of the PWN accompanied by PSR B1509$-$58 with 190 ks exposure taken by {\\itshape Chandra}. The image was smoothed with a Gaussian of $\\sigma=2.0''$. The white rectangle indicates the FoV of the right panel. (Right)--- Close-up image of the pulsar vicinity smoothed with a Gaussian of $\\sigma=0.5''$. The inner panel schematically describes the remarkable structures.} \\label{X-ray-image} \\end{figure*} So far, the shock front of the pulsar wind has been observed only in the Crab nebula \\citep{2000ApJ...536L..81W} because of the small angular sizes of PWNe. Nevertheless the {\\itshape Chandra} X-ray observatory with sub-arcsec-scale spatial resolution has the potential to detect the wind termination shocks in other PWNe. In order to obtain a general understanding of the formation process of PWNe, we investigate the fine structures around the pulsar PSR B1509$-$58 utilizing the {\\itshape Chandra} data. PSR B1509$-$58 is emitting 150 ms radio, X-ray, and Gamma-ray pulsation \\citep{1982ApJ...256L..45S,1982ApJ...262L..31M,1993ApJ...417..738U}. From the spin parameters, a characteristic age $\\tau_\\mathrm{c}=1700$ yr, a spin-down luminosity $\\dot{E} = 1.8 \\times 10^{37}$ ergs s$^{-1}$, and a surface magnetic field $B_\\mathrm{p}=1.5\\times10^{13}$ G have been obtained, making it one of the youngest, the most energetic, and highest field pulsars known \\citep{1982ApJ...256L..45S,1994ApJ...422L..83K}. The pulsar is accompanied by a bright and large PWN with remarkable jet-like features extending to the southeast and northwest \\citep{2002ApJ...569..878G}. The pulsar and its PWN appears to be embedded in a 30 arcmin radio shell G320.4$-$01.2 \\citep{1981MNRAS.195...89C}. An H{\\rm I} observation yielded the distance of $d=5.2\\pm1.4$ kpc from the earth \\citep{1999MNRAS.305..724G}. Adopting this distance and standard parameters of the ISM and supernova explosions yields an age of 6-20 kyr, which is one order of magnitude larger than the pulsar's characteristic age \\citep{1983ApJ...267..698S}. Moreover \\citet{2005XrU...604..379Y} detected proper motions of X-ray clumps in RCW 89 which coincide with the north radio shell from the 4.3 yr base-line {\\itshape Chandra} observations. These results imply that the pulsar and the SNR have the same progenitor in a cavity. If this is the case the termination shock of the pulsar wind can be easily discriminated by {\\itshape Chandra}. The paper is structured as follows. The {\\itshape Chandra} observations and their results are shown in \\S2. The imaging analysis and spectral analysis are described in \\S3 and \\S4, respectively. Finally, we discuss the obtained results in \\S5. ", "conclusions": "\\subsection{Double tori} As shown in Figure \\ref{radial-profile}, the PWN around PSR B1509$-$58 has a centrally concentrated surface brightness profile, which may be caused by its comparatively small angular size or the contamination from the bright jet structure pointing towards us. For this reason, it is difficult to compare the surface brightness evolution of PSR B1509$-$58 with that of the Crab nebula in a simple way. Nevertheless PSR B1509$-$58 possesses an apparently different morphology; the nested double ring structure, inner-torus ($R\\sim30''$) and outer-torus ($30'' 200$ pc. Since the low latitude contribution as a whole is likely to be inflated by the contribution of distant emission, this figure is a lower limit. \\\\\\par \\subsection{Remarks on Mass Estimation} It is worth making a few comments on the uncertainties associated with molecular cloud mass estimation, which are often passed over in CO cloud studies. The empirical `X-factor', used to convert $^{12}$CO luminosity to H$_2$ column density, has been derived from a number of independent methods, and is believed to be accurate to a factor of around two, when averaged over large cloud populations \\citep[see e.g.][]{dame01, hunter97, solomon91}. For individual clouds the uncertainty will be larger \\citep[e.g.][]{magnani95}. In particular, \\citet{heyer01} have demonstrated how the correct X-factor for a cloud depends on the gravitational state of the gas, and argue that it may significantly overestimate the masses of objects that are not bound by self-gravity. Conversely, several authors have used the results of PDR modeling to argue that the standard Galactic X-factor may significantly \\textit{underestimate} the masses of many molecular clouds \\citep{bell06, kaufman99, pringle01}, especially those that are particularly diffuse. $^{13}$CO LTE-based mass estimates are also subject to their own uncertainties. \\citet{padoan00} use model clouds with realistic temperature and density structures to argue that LTE analysis may underestimate true column density by a factor of 1.3 to 7, with the most extreme differences occurring in regions of low column density. The LTE values of $N(^{13}\\mathrm{CO})$ observed in the present dataset range from $\\sim5\\times10^{14}$ cm$^{-2}$ to $\\sim8\\times10^{15}$ cm$^{-2}$, which corresponds to a factor of $N(^{13}\\mathrm{CO})_{true}/N(^{13}\\mathrm{CO})_{LTE}~\\lesssim~3.5$ in their analysis. The unavoidable assumption of a constant H$_2$ to $^{13}$CO abundance ratio may also introduce error into the results. CO catalogues such as the one presented here offer an attractive means of calibrating molecular cloud mass conversion factors, in particular the X-factor, through comparison with data from the next generation of gamma ray telescopes such as GLAST. Molecular clouds act as passive targets for hadronic cosmic ray interactions, which result in the production of gamma rays in the MeV to TeV regime. The mass of the target cloud may be calculated directly from the gamma-ray luminosity, and this mass may be compared to CO luminosity without the need for many of the assumptions that plague other calibration methods. Unfortunately, the estimated masses and angular extents of the catalogue clouds place most of them well below the detectability threshold for GLAST \\citep{torres05}. Nevertheless, several high-latitude objects -- notably clouds 74 and 16 -- fall on the border of predicted range and may eventually prove to be detectable. In the near future these clouds, and others like them, may offer the opportunity to calibrate the X-factor for high altitude clouds away from the immediate solar neighborhood." }, "0808/0808.3582_arXiv.txt": { "abstract": "We present the first detection and mapping of the HD 32297 debris disk at 1.3 mm with the Combined Array for Research in Millimeter-wave Astronomy (CARMA). With a sub-arcsecond beam, this detection represents the highest angular resolution (sub)mm debris disk observation made to date. Our model fits to the spectral energy distribution from the CARMA flux and new Spitzer MIPS photometry support the earlier suggestion that at least two, possibly three, distinct grain populations are traced by the current data. The observed millimeter map shows an asymmetry between the northeast and southwest disk lobes, suggesting large grains may be trapped in resonance with an unseen exoplanet. Alternatively, the observed morphology could result from the recent breakup of a massive planetesimal. A similar-scale asymmetry is also observed in scattered light but not in the mid-infrared. This contrast between asymmetry at short and long wavelengths and symmetry at intermediate wavelengths is in qualitative agreement with predictions of resonant debris disk models. With resolved observations in several bands spanning over three decades in wavelength, HD 32297 provides a unique testbed for theories of grain and planetary dynamics, and could potentially provide strong multi-wavelength evidence for an exoplanetary system. ", "introduction": "Debris disks provide the principal means of studying the formation and evolution of planetary systems on timescales of 10$-$100 Myr. Evidence for exoplanets in these systems can be found by matching density variations in debris disks to theoretical models of the gravitational perturbations caused by planets (e.g., Reche et al. 2008). A modest sample of debris disks have now been imaged in the visible, and many show substructure such as clumps, warps, and offsets, consistent with dynamical perturbations by massive planets. However, a wide variety of other mechanisms can produce similar structures (Moro-Martin et al. 2007, and references therein). As different wavebands are sensitive to different grain sizes, which are in turn subject to different dynamical influences, multi-wavelength observations offer the most promising path towards definitively classifying the physical mechanisms at work in these systems. At present, only a few debris disks have resolved observations spanning more than a decade in wavelength. A particularly critical, though technologically challenging, deficit of observations lies at (sub)millimeter wavelengths, which trace large grains primarily affected by gravitational forces. To date, bolometer arrays have resolved 8 debris disks in the (sub)mm (e.g., Holland et al. 1998, Greaves et al. 1998). However, such low-resolution ($\\theta_{\\mathrm{beam}} \\gtrsim 10''$) single-dish measurements are limited to the largest, nearest disks. Higher resolution interferometric observations are needed to access the larger debris disk population already imaged at shorter wavelengths. Some pioneering work has been done in this area; OVRO and PdBI have detected and resolved two debris disks (Vega: Koerner et al. 2001, Wilner et al. 2002; HD 107146: Carpenter et al. 2005). Recently, \\citet{Corder08} mapped HD 107146 with the Combined Array for Research in Millimeter-wave Astronomy (CARMA) at 1.3 mm, providing the highest fidelity interferometric debris disk map to date. Here, we report the near-simultaneous CARMA detection of HD 32297, the third debris disk mapped with a (sub)mm interferometer. With a sub-arcsecond beam, this detection is the highest angular resolution (sub)mm debris disk observation made to date. HD 32297 is a $\\sim$30 Myr A-star at $112_{-12}^{+15}$ pc \\citep{Perryman97}, first discovered to host a resolved debris disk with HST/NICMOS near-infrared (NIR) imaging \\citep{Schneider05}. The discovery image showed an edge-on debris disk extending to 400 AU (3.3$''$), with an inner-disk brightness asymmetry inward of 60 AU ($0.5''$). \\citet{Kalas05} subsequently imaged HD 32297 in the optical, revealing an asymmetric, extended outer disk ($\\sim$1700 AU, 15$''$) likely interacting with the interstellar medium. Later, \\citet{Redfield07} detected circumstellar gas in this system, reporting the strongest Na I absorption measured toward any known debris disk. Most recently, \\citet{Fitzgerald07}, hereafter F07, and \\citet{Moerchen07} resolved HD 32297 in mid-infrared (MIR), thermal emission. Detailed analysis of the spectral energy distribution (SED) by F07 showed that multiple grain populations may be present in the disk. The lack of long wavelength data needed to characterize the large grain properties of HD 32297 motivated the CARMA observations presented here. ", "conclusions": "F07 attempted to model the observed SED and $N'$-band image of HD 32297 with a single ring of grains of characteristic size, but found that a second population of grains was needed to adequately fit the observed SED for $\\lambda \\gtrsim 25 \\mu$m. Indeed, the contrast between the observed mm and $N'$-band morphologies (see below and Figure \\ref{multiwav}) suggests that the grains responsible for the emission at each wavelength constitute separate populations. To test whether the population of mm-emitting grains is consistent with the second, larger grain population proposed by F07 to fit the SED at 25 $\\mu$m $\\lesssim \\lambda \\lesssim$ 60 $\\mu$m, we revisit the F07 model, adopting their data and fitting method, and incorporating the Qa-band flux of \\citet{Moerchen07} and the new MIPS and CARMA fluxes. The free parameters in the single population model of F07 are the disk inclination ($i$), position angle (PA), radii of the inner and outer edges ($\\varpi_0$, $\\varpi_1$), surface density power-law index ($\\gamma$), vertical optical depth to absorption at the inner edge ($\\tau_0\\equiv\\tau_\\perp^{\\mathrm{abs}}(\\varpi_0)$), stellar flux factor ($\\xi$), and effective grain size ($\\lambda_\\mathrm{sm}$). To fit the long-wavelength SED ($\\lambda \\gtrsim 25 \\mu$m), we augment this model with a population of larger grains of effective size $\\lambda_\\mathrm{lg}$ and total emitting area $A_\\mathrm{lg}$, located in a narrow ring at the small-grain inner-disk edge, $\\varpi_0$. These grains contribute flux, $F_{\\nu,\\mathrm{lg}}$, according to: \\begin{eqnarray} \\epsilon_{\\nu,\\mathrm{lg}} &=&\\left\\{ \\begin{array}{ll} \\lambda_\\mathrm{lg}/\\lambda & \\mbox{if $\\lambda>\\lambda_\\mathrm{lg}$,} \\\\ 1 & \\mbox{otherwise}, \\end{array}\\right. \\\\ T_\\mathrm{lg}(r) &=& 468 \\left(\\frac{L_*/L_\\sun}{\\lambda_\\mathrm{lg}/1\\,\\micron}\\right)^{1/5} \\left(\\frac{r}{1\\,\\mathrm{AU}}\\right)^{-2/5} \\mathrm{K}, \\\\ F_{\\nu,\\mathrm{lg}} &=& \\left(\\frac{A_\\mathrm{lg}}{d^2}\\right) \\epsilon_{\\nu,\\mathrm{lg}} B_\\nu[T_\\mathrm{lg}(\\varpi_0)]. \\end{eqnarray} Following F07, we ran three Monte Carlo Markov chains with $3\\times 10^4$ samples, and simulataneously fit both the large and small grain populations. The results of this procedure are listed in Table \\ref{sed_table}, and the corresponding range of allowed dust emission is plotted in Figure \\ref{sed} (red: small grains, blue: large grains, grey: photosphere, green: composite). While the two-population model provides a satisfactory fit to the mid- and far-infrared data, the mm flux is underestimated at the 4$\\sigma$ level, suggesting that the second grain population proposed by F07 is not responsible for the majority of the mm flux. Nevertheless, the new MIPS data lend further evidence to the F07 suggestion that two populations are needed to fit the observed SED for $\\lambda \\lesssim 160 \\mu$m. Therefore, the fit suggests at least three distinct populations are traced by the current observations. However, from the two-population model, only the 1.3 mm flux appears to trace the putative third population. Since this population is described by both a mass/emitting area and a size/temperature, we lack sufficient data to fully characterize it via modeling. We note, though, that this population likely traces $\\gtrsim 95$\\% of the total dust mass. Assuming a characteristic stellocentric distance of $\\sim$50 AU (\\S 3), $L_{*}=5.4L_{\\sun}$ (F07), and an effective grain size of 1.3-mm, the implied mm grain temperature is $\\sim$30 K, suggesting a dust mass of $M_{\\mathrm{mm}}\\sim M_{\\earth}$ (adopting an opacity of 1.7 cm$^2$ g$^{-1}$). This estimated mass is among the highest observed for debris disks detected in the (sub)mm and is two orders of magnitude larger than that implied for the ``large-grain'' population in the SED fit: $M_{\\mathrm{lg}}\\sim0.02M_{\\earth}$ (using the fitted parameters in Table \\ref{sed_table} and assuming spherical grains with a density of 1 g cm$^{-3}$). Future far-infrared / sub-mm observations are needed to confirm the three populations proposed here and better constrain their properties. \\begin{deluxetable}{lcc} \\tablecaption{\\label{sed_table} Best-Fit Model Parameters} \\tablewidth{0pt} \\tablehead{ \\colhead{Parameter} & \\colhead{Best-fit} & \\colhead{Description} } \\startdata $i$ (deg) & $90 \\pm 5$ & disk inclination \\\\ PA (deg) & $46 \\pm 3$ & disk position angle \\\\ $\\varpi_0$ (AU) & $70_{-10}^{+20}$ & inner edge \\\\ $\\varpi_1$ (AU) & $>$ 1200 & outer edge \\\\ $\\log_{10}\\tau_0$ & $-2.4_{-0.2}^{+0.3}$ & vertical optical depth at $\\varpi_0$ \\\\ $\\gamma$ & $<$ -1.63 & surf density power-law index \\\\ $\\xi / 7.22 \\times 10^{-20}$ & $1.03 \\pm 0.03$ & $(R_{*}/d)^2$, stellar flux factor \\\\ $\\log_{10}(\\lambda_\\mathrm{sm}/1\\,\\micron)$ & $-1.4_{-1.3}^{+0.6}$ & small grain effective size \\\\ $\\log_{10}(A_\\mathrm{lg}/1\\,\\mathrm{cm}^2)$ & $28.9 \\pm 0.2$ & large grain emitting area \\\\ $\\log_{10}(\\lambda_\\mathrm{lg}/1\\,\\micron)$ & $1.3_{-0.1}^{+0.2}$ & large grain effective size \\enddata \\tablecomments{Confidence intervals are 95\\% for marginal posterior distributions. Adopted priors are as in F07, except for $\\gamma$, which is constrained to be in the domain [-4,0] due to the inconsistency of rising surface density with the scattered-light image and CARMA map; the large grain effective size was constrained by a log-uniform prior from 1\\,nm to 100\\,mm and a requirement that $\\lambda_{\\mathrm{lg}} > \\lambda_{\\mathrm{sm}}$.} \\end{deluxetable} \\begin{figure} \\centering \\includegraphics[width=0.23\\textwidth,angle=90]{f4.eps} \\caption{{\\it Left:} CARMA contours of HD 32297, overlaid on the NIR scattered-light image from Schneider et al. (2005). {\\it Right:} Photosphere-subtracted MIR contours from F07 overlaid on the same image. The asymmetry in the CARMA data between the northeast and southwest lobes suggests the large, mm-sized grains may be trapped in resonance with an unseen exoplanet. A similar asymmetry is also observed in scattered light but not in the MIR. The contrast between asymmetry at short and long wavelengths and symmetry at intermediate wavelengths is a direct prediction of the resonant debris disk models of Wyatt (2006).} \\label{multiwav} \\end{figure} To address the observed mm morphology, in Figure \\ref{multiwav}, we qualitatively compare the CARMA mm map (contours, left panel) to the MIR image of F07 (contours, right panel) and the NIR scattered-light image of \\citet{Schneider05} (color, both panels). The images at all three wavelengths are consistent with an edge-on disk. However, while both the NIR and CARMA data exhibit a brightness asymmetry between the northeast and southwest lobes inward of $\\sim$0.5$''$, the F07 MIR image is consistent with azimuthal symmetry. We note that \\citet{Moerchen07} found evidence for asymmetry in their Qa-band data, though their data were not PSF subtracted, and the asymmetry was in the opposite sense as observed in the NIR / CARMA data (NE lobe brighter than SW). Recently, \\citet{Grigorieva07} used predictions from their numerical model of collisional avalanches to suggest that the observed scattered-light asymmetry in HD 32297 results from the breakup of a large planetesimal. However, while a massive collision can explain the observed NIR and mm morphology, it is not clear why the MIR image would not show a similar asymmetry. An alternative hypothesis is that the structure results from a planetary-induced resonance. In this case, the dust is either trapped in resonance as it drifts inward due to Poynting-Robertson drag, or it remains locked in resonance after being generated by parent planetesimals in resonance themselves (e.g., Krivov et al. 2007, and references therein). Interestingly, a recent study of the latter mechanism by \\citet{Wyatt06} directly predicts a contrast between asymmetry at short and long wavelengths and symmetry at intermediate wavelengths, as observed in HD 32297. In this model, (sub)mm emission is dominated by large grains, which have the same clumpy resonant distribution as the parent planetesimals. Small grains, traced at short wavelengths, exhibit a similar asymmetry, as they are preferentially born in the high density, resonant structures before being rapidly expelled from the system. Lastly, moderately-sized grains sampled at intermediate wavelengths remain bound to the star, but have fallen out of resonance due to radiation pressure and are subsequently scattered into an axisymmetric morphology. This proposed scenario of \\citet{Wyatt06} conveniently explains the qualitative picture for HD 32297 depicted in Figure \\ref{multiwav}, yet we caution that this suggestion is highly speculative, and future rigorous modeling of this system is needed to draw firm conclusions from the available data. One potential problem with this hypothesis is that the SED model predicts that the small ($N'$-band-emitting) grains have sizes $\\lesssim$ 1 $\\mu$m (Table \\ref{sed_table}), similar to that expected to produce the NIR scattered-light image (see discussion in F07). However, the interpretation of Figure \\ref{multiwav} in terms of the \\citet{Wyatt06} models requires that the NIR-scattering and MIR-emitting grains have different sizes. This ambiguity could potentially be resolved through simultaneous modeling of scattering and emission, incorporating the optical image presented in \\citet{Kalas05}. Most importantly, and independent of speculation, HD 32297 is currently one of only a few debris disks with resolved observations in four wavelength regimes (optical, NIR, MIR, mm). Taken together, these observations can provide a unique testbed for theories of grain and planetary dynamics." }, "0808/0808.1855_arXiv.txt": { "abstract": "The microscopic quantum field theory origins of warm inflation dynamics are reviewed. The warm inflation scenario is first described along with its results, predictions and comparison with the standard cold inflation scenario. The basics of thermal field theory required in the study of warm inflation are discussed. Quantum field theory real time calculations at finite temperature are then presented and the derivation of dissipation and stochastic fluctuations are shown from a general perspective. Specific results are given of dissipation coefficients for a variety of quantum field theory interaction structures relevant to warm inflation, in a form that can readily be used by model builders. Different particle physics models realising warm inflation are presented along with their observational predictions. ", "introduction": "It has been fourteen years since warm inflation was introduced and with it the first and still only alternative dynamical realisation of inflation to the standard scenario. The standard picture of inflation introduced in 1981 relied on a scalar field, called the inflaton, which during inflation was assumed to have no interaction with any other fields. During inflation, this field rolls down its potential and due to it being coupled to the background metric, a damping-like term is present which slows down its motion. As this inflaton field was assumed to not interact with other fields, there was no possibility for radiation to be produced during inflation, thus leading to a thermodynamically supercooled phase of the Universe during inflation. Getting out of this inflation phase and putting the Universe into a radiation dominated phase was a key issue, termed the ``graceful exit\" problem \\cite{oldinf,ni,reheatu}. The first successful solution of the graceful exit problem and so the first successful cold inflation model, was new inflation \\cite{ni}. The solution was to picture particle production as a distinct separate stage after inflationary expansion in a period called reheating. In this phase, couplings to other fields were assumed to be present and the inflaton would find itself in a very steep potential well in which it would oscillate. These oscillations would lead to a radiative production of particles. There have been many variants of the original cold inflation picture, first introduced in the context of the new inflation model and shortly afterwards in the chaotic inflation model~\\cite{ci}, with many other models that followed. The warm inflation picture differs from the cold inflation picture in that there is no separate reheating phase in the former, and rather radiation production occurs concurrently with inflationary expansion. The constraints by General Relativity for realising an inflationary phase simply require that the vacuum energy density dominates and so this does not rule out the possibility that there is still a substantial radiation energy density present during inflation. Thus on basic principles, the most general picture of inflation accommodates a radiation energy density component. The presence of radiation during inflation implies the inflationary phase could smoothly end into a radiation dominated phase without a distinctively separate reheating phase, by the simple process of the vacuum energy falling faster than the radiation energy, so that at some point a smooth crossover occurs. This is the warm inflation solution to the graceful exit problem. Dynamically warm inflation is realised if the inflaton were interacting with other fields during the inflation phase. In fact in any realistic model of inflation, the inflaton must be coupled to other fields, since eventually the inflaton must release its vacuum energy to other fields thereby creating particles which form the subsequent radiation dominated era in the Universe. Thus the idea that these couplings to other fields somehow are inactive during inflation, as pictured in the cold inflation picture, is something that does require verifying by detailed calculation. When such calculations are done, the result is that there are regimes in which particle production during inflation occurs. This review will present the calculations which demonstrate particle production during inflation, thereby leading to a warm inflationary expansion. The idea of particle production concurrent with inflationary expansion was first suggested in the pre-inflation inflation paper by L.Z. Fang in 1980 \\cite{Fang:1980wi}. His paper proposed using a scalar field with the origin of inflationary expansion due to a claimed anomalous dissipation term that would be generated based on Landau theory if the field was undergoing a second order phase transition. This was dynamically very different from the scalar field inflation that eventually became successful, and the source of dissipation was also different from that in warm inflation. However this model captured the basic idea of concurrent particle production and inflation. Then in the mid-80s two papers proposed adding a local $\\Upsilon {\\dot \\phi}$ type dissipation term into the evolution equation of the inflaton field, Moss \\cite{im} and then Yokoyama and Maeda~\\cite{Yokoyama:1987an}. In both cases the dissipative term generated a source of radiation production during inflation. The idea of a dissipative term was re-discovered independently by Berera and Fang~\\cite{Berera:1995wh} almost a decade later. They went further by proposing that the consistent dynamics of the inflaton field was a Langevin equation, in which a fluctuation-dissipation theorem would uniquely specify the fluctuations of the inflaton field. That paper by Berera and Fang provided the foundations for the theory of fluctuations in warm inflation and the Langevin equation has since been the fundamental equation governing inflaton dynamics. {}Following that work, in~\\cite{Berera:1995ie} Berera proposed that a separate reheating phase, as standard in all inflation models up to then, could be eliminated altogether. This paper proposed a new picture of inflation, which it termed warm inflation, in which the process of inflationary expansion with concurrent radiation production could terminate simply by the radiation energy over-taking the vacuum energy, thus going from an inflationary to a radiation dominated era. This work presented an alternative solution to the graceful exit problem to the one given by the standard inflation scenario. This warm inflation picture was verified explicitly in \\cite{Berera:1996fm}, where the {}Friedmann equations for a Universe consisting of vacuum and radiation energy were studied and gave many exact warm inflation solutions to the graceful exit problem. {}Finally in~\\cite{Berera:1999ws} the calculation of fluctuations were done by Berera for the inflaton evolving by a Langevin equation in a thermal inflationary Universe. Alongside the development of the basic scenario, the first principles quantum field theory dynamics of warm inflation was developed. This started in \\cite{Berera:1996nv} with a quantum mechanical model which demonstrated the origin of the fluctuation-dissipation relation in warm inflation. The key step in deriving warm inflation from quantum field theory is in realising an overdamped regime for the evolution of the background inflaton field. The initial attempt at this was done by Berera, Gleiser and Ramos in~\\cite{BGR}. In this work, it was proposed that the overdamped evolution should occur under adiabatic conditions in which the microscopic dynamical processes operated much faster than all macroscopic evolution, in particular the scalar field motion and Hubble expansion. Based on this criteria, a set of consistency conditions were formulated in \\cite{BGR}, which would be required for a self-consistent solution. However in~\\cite{BGR} no explicit warm inflation solutions were found. This point was further highlighted by Yokoyama and Linde in~\\cite{Yokoyama:1998ju}, in which several models were studied from which the conclusions of~\\cite{BGR} were verified. The problem in these early works was that dissipation effects were being looked at in a high temperature regime and it proved too difficult to keep finite temperature effective potential corrections small, so that the inflaton potential remained relatively flat, and at the same time obtain a large dissipative coefficient. One type of model was shown that could realise such requirements~\\cite{BGR2} and this was the first quantum field theory model of warm inflation, although it was not a very compelling model. Subsequently Berera and Ramos in \\cite{BR1} suggested a solution for getting around the mutual constraints of obtaining a large dissipative coefficient and yet small effective potential corrections. The main observation was that supersymmetry can cancel local quantum corrections, such as zero temperature corrections to the effective potential, whereas temporally non-local quantum effects, such as those that underly the dissipative effects, will not be cancelled. This led to \\cite{BR1} proposing a two-stage interaction configuration, in which the inflaton was coupled to heavy ``catalyst'' fields with masses larger than the temperature of the Universe and these fields in turn were coupled to light fields. The evolution of the inflaton would induce light particle production via the heavy catalyst fields. Since these heavy catalyst fields were basically in their ground state, the quantum corrections associated with them could be cancelled in supersymmetric models. The calculation of the low temperature dissipative coefficients for this two-stage mechanism were first done by Moss and Xiong \\cite{mx}. There is an earlier review which covered the basics of the warm inflation scenario \\cite{Berera:2006xq}. In this review full details will be developed of the quantum field theory dynamics of warm inflation. This will first start in Sec. \\ref{wipicture} with a summary of the warm inflation scenario, including a comparison of it to cold inflation. In Sec. \\ref{sec TFT} a basic introduction to thermal field theory is given including the real time formalism for interacting field theories. In Sec. \\ref{effeom} the effective evolution equation of the inflaton field is derived, in which all fields it interacts with are integrated out, leading to a Langevin type nonconservative equation which contains a dissipative term and a noise force term. The detailed properties of these dissipation and fluctuation terms is then studied in Sec. \\ref{fd}. In addition the physical picture of the dissipation effects in warm inflation are discussed. In Sec. \\ref{FRWspacetime} the calculations are extended to curved spacetime. In Sec. \\ref{particlemod} various particle physics models of warm inflation are presented. Finally Sec. \\ref{concl} presents some concluding remarks and future work being done on the subject. Our conventions are as follows. We use spacetime metrics with $(+---)$ signature, and we use natural units for the Planck's constant, Boltzmann's constant and the velocity of light $\\hbar=k=c=1$. ", "conclusions": "\\label{concl} Generically, the inflaton interacts with other fields in any typical inflation model and so its dynamics is dissipative. As such, inflation, like most dynamics in nature is an open system phenomenon, thus requiring a much more complex analysis of its dynamics than the one typically formulated for the cold inflation picture. This general point has been voiced by B. L. Hu and coworkers ~\\cite{hu1} and more specifically in the initial motivating papers of warm inflation \\cite{Berera:1995wh,Berera:1995ie,Berera:1996nv,BGR}. In all these cases, the point has been made that the problem of inflation encompasses many different branches of Physics, from nonequilibrium statistical dynamics to particle physics phenomenology. Warm inflation dynamics is a rich area of study for both quantum field theory real time dynamics and for particle physics model building. The study of the warm inflation dynamics has almost exclusively motivated the understanding of strong dissipative behavior in quantum field theory. This started initially with the work of Berera, Gleiser and Ramos~\\cite{BGR}, in which dissipation was examined in the overdamped regime for the first time in quantum field theory using extended linear response calculations. Since then, specific progress has been made to underpin interaction structures in interacting quantum field theory models which lead to strong dissipative behavior under warm inflationary conditions~\\cite{BR1,BRplb1,BRfrw,BRplb2}. At a more general level, this work has motivated the first calculations of dissipation in the low temperature regime~\\cite{mx}. In this review, a physical picture has also been developed in Sec.~\\ref{fd}, for explaining the dissipation behavior found in warm inflation from the quantum field theory calculations, and this is further developed in~\\cite{grahammoss}. Up to now all these studies has been based on variants of linear response methods, in which case the dynamics is supposed to happen close to equilibrium, or in a quasi-adiabatic regime, including various types of resummations. Several extensions to this work, which will give a more accurate understanding of dissipation, are under way. {}For instance, when departing from the quasi-adiabatic regime for the field dynamics, it is expected that the local Markovian approximation commonly used to analyse the field equations can differ significantly from the exact nonlocal equations. Preliminary tests have shown that this difference can be very large for a period of time starting from the initial period, but with differences between the dynamics getting smaller at longer times, depending on the model parameters~\\cite{FRS}. The methods used in this review are all based on the effective equation of motion derived from the action functional. In order to study strong nonequilibrium dynamics, it requires methods beyond these. In this case, full kinetic set of equations for the relevant fields have to be studied, which contain information not only of the inflaton effective dynamics but also about thermalisation and equilibration. Work in this direction, also can help to better understand the physics of particle production during the system dynamics and be used to check the reliability of using the approximation of thermal initial conditions for the bath fields. Initial work in this direction, though not directly related to the type of models relevant to warm inflation as discussed extensively e.g. in Sec.~\\ref{particlemod}, has appeared~\\cite{Aarts:2007ye}, while a work more on the warm inflation dynamics motivated side is under way~\\cite{BMR}. This review has presented in detail the methods used so far to understand warm inflation dynamics as well as the limitations of these studies. In particular these methods used so far are primarily quasi-adiabatic approximations for the fields with the assumption of near thermal equilibrium evolution. Despite these limitations, these results find parameter regimes in which physically acceptable solutions exist over time periods sufficiently long to be of use in studies of warm inflation. During this time interval in which these approximations apply, and where a local Markovian dynamics can be used, in contrast to the full non Markovian one, in the typical model implementations discussed in Sec.~\\ref{particlemod}, the amount of radiation production was seen to be sufficient to change the cold inflationary picture predictions regarding the density perturbations, thus requiring its description in terms of the warm inflation picture. Moreover, once dissipative effects are strong enough, inflation can be sustained and driven longer than when these effects are neglected. Consequently, parameter values typically required in the cold inflation case can be relaxed, which can help to evade various problems that plaque the standard scenario of inflation, like the graceful exit and $\\eta$-problem, discussed in details in this review, as well as the problems of quantum-to-classical transition~\\cite{Berera:1995ie,bel} and the initial conditions for inflation~\\cite{BG,ROR}. The development of the quantum field theory dynamics of warm inflation has in turn been applied to particle physics model building, in which warm inflation dynamics in realised. Early on it was recognisied in \\cite{Berera:1999ws} that warm inflation has some appealing and unique model building features. In particular, it offers a simple solution to the $\\eta$-problem and for monomial potentials, observationally consistent inflation can occur for the inflaton amplitude below the Planck scale $\\langle \\phi \\rangle < m_P$. These features have been realised in explicit first principles quantum field theory models of warm inflation in \\cite{BGR2,Berera:1999ws,bb4,BuenoSanchez:2008nc}. These studies are now being extended to develop a complete particle cosmology in which not only is warm inflation realised, but in addition other features such as leptogenesis and gravitino abundances are addressed, developing in depth some of the work already started on these topics \\cite{Taylor:2000jw,bb1,bb4,Lambiase:2006md,BuenoSanchez:2008nc}. In a separate direction, the warm inflation models developed so far in \\cite{bb4,BuenoSanchez:2008nc} have been in the low temperature regime. In \\cite{bbrecent} this is being extended to higher temperatures. This requires calculating all the thermal loop corrections in the SUSY models that have the two-stage interaction structure Eq. (\\ref{wi2stage}) relevant for warm inflation. Some initial work has been done in \\cite{Hall:2004zr}, and a more detailed analysis is now underway \\cite{bbrecent}." }, "0808/0808.0455_arXiv.txt": { "abstract": "We present new \\XMM EPIC observations of the nuclei of the nearby radio galaxies 3C\\,305, DA\\,240, and 4C\\,73.08, and investigate the origin of their nuclear X-ray emission. The nuclei of the three sources appear to have different relative contributions of accretion- and jet-related X-ray emission, as expected based on earlier work. The X-ray spectrum of the FRII narrow-line radio galaxy (NLRG) 4C\\,73.08 is modeled with the sum of a heavily absorbed power law that we interpret to be associated with a luminous accretion disk and circumnuclear obscuring structure, and an unabsorbed power law that originates in an unresolved jet. This behavior is consistent with other narrow-line radio galaxies. The X-ray emission of the low-excitation FRII radio galaxy DA\\,240 is best modeled as an unabsorbed power law that we associate with a parsec-scale jet, similar to other low-excitation sources that we have studied previously. However, the X-ray nucleus of the narrow-line radio galaxy 3C\\,305 shows no evidence for the heavily absorbed X-ray emission that has been found in other NLRGs. It is possible that the nuclear optical spectrum in 3C\\,305 is intrinsically weak-lined, with the strong emission arising from extended regions that indicate the presence of jet--environment interactions. Our observations of 3C\\,305 suggest that this source is more closely related to other weak-lined radio galaxies. This ambiguity could extend to other sources currently classified as NLRGs. We also present \\XMM and VLA observations of the hotspot of DA\\,240, arguing that this is another detection of X-ray synchrotron emission from a low-luminosity hotspot. ", "introduction": "\\label{intro} Radio galaxies consist of twin jets of particles that are ejected from a compact region in the vicinity of a supermassive black hole, feeding into large-scale `plumes' or `lobes'. There are two principal morphological classes of radio galaxies, low-power (Fanaroff-Riley type I, hereafter FRI) sources and high-power (FRII) sources \\cite{fr74}. FRI sources exhibit `edge-darkened' large-scale radio structure, and modeling implies that initially supersonic jets in these sources decelerate to transonic speeds on $\\sim$kpc scales before flaring into large plumes (e.g., \\citealt{per07}). FRII sources appear `edge-brightened', and in these cases highly supersonic jets propagate out to large distances (often $>100$ kpc) from the core before terminating in bright hotspots and accompanying radio lobes. Observationally, the Fanaroff-Riley divide occurs at a 178-MHz radio power of $\\sim$$10^{25}$ W~Hz$^{-1}$~sr$^{-1}$. It is important to understand whether the kpc-scale Fanaroff-Riley dichotomy is determined by the interaction between the jet and its external hot-gas environment (e.g., \\citealt{bic95}), or rather is nuclear in origin and governed by differences in the properties of the accretion flow (\\citealt{rey96}). The first observations of large samples of $z<0.1$ 3CRR radio-galaxy {\\it nuclei} with \\Ch and \\XMM (\\citealt{don04,bal06,evans06}) showed that FRI nuclei show no signs of heavily absorbed X-ray emission that would be expected from standard AGN unification models (\\citealt{up95}), are dominated by emission from an unresolved jet (e.g.,~\\citealt{wb94,bal06,evans06}), and have highly radiatively inefficient accretion flows. Narrow optical-line FRII sources show evidence for heavily obscured ($N_{\\rm H}>10^{23}$~cm$^{-2}$) nuclear X-ray emission that is associated with a radiatively efficient accretion flow, together with an unabsorbed component of jet-related emission (\\citealt{evans06,hec06}). FRII radio galaxies at higher redshift are consistent with such behavior (\\citealt{bel06}). A significant breakthrough for understanding the physical origin of the FRI/FRII dichotomy came from \\Ch and \\XMM observations of the population of {\\it low-excitation radio galaxies} (LERGs), which have weak or no emission lines in their optical spectra (\\citealt{hine79,jac97}). Almost all FRI radio galaxies are LERGs, but there is a significant population of FRII LERGs at $0.1 < z < 0.5$. The X-ray spectra of LERGs, irrespective of their FRI or FRII morphology, are dominated by unabsorbed emission that can be associated with a parsec-scale jet, with no obvious contribution from accretion-related emission. These sources are likely to accrete in a radiatively inefficient manner (\\citealt{hec06}). On the other hand, high-excitation radio galaxies (HERGs -- i.e., NLRGs, BLRGs, and quasars), which display prominent narrow or broad optical emission lines, have X-ray spectra that are consistent with standard unification models: they show evidence for luminous, radiatively efficient accretion disks, together with circumnuclear tori when the source is oriented close to edge-on with respect to the observer. HERGs tend to show evidence for additional hot dust over and above that of LERGs in their mid-IR spectra (e.g., \\citealt{ogle06}; Birkinshaw et al., in preparation), which is again consistent with reprocessing of luminous accretion-related emission by torus-like structure. Most high radio-power (FRII) sources are high-excitation radio galaxies. The distinct X-ray nuclear properties of low- and high-excitation radio galaxies, regardless of their large-scale FRI or FRII morphology, could be interpreted as implying that the Fanaroff-Riley dichotomy remains principally influenced by jet power and environment. The excitation dichotomy, on the other hand, is interpreted to be attributed to the radiative efficiency of the accretion flow (e.g., \\citealt{hec06}) and possibly related to the nature of the accreting material (\\citealt{hec07}). Here, we report new \\XMM observations of the nuclei of three $z<0.1$ 3CRR radio galaxies --- 3C\\,305, DA\\,240, and 4C\\,73.08. The three sources have 178-MHz radio powers that lie close to the FRI/FRII dividing luminosity (Table~\\ref{sourcessummary}), plus a range of radio morphologies and optical emission-line characteristics. They are therefore good candidates for examining possible connections between the central engine and large-scale radio characteristics. This paper is organized as follows. In Section 2, we describe the optical and radio properties of the three sources. Section 3 contains a description of the data and a summary of our analysis. In Section 4, we report the results of our spectroscopic analysis of the sources. In Section 5, we describe VLA and \\XMM observations of the bright NE hotspot in DA\\,240. In Section 6, we interpret the observations in the context of our previous \\Ch and \\XMM observations of 3CRR radio galaxies and discuss the optical emission-line characteristics of the sources. We end with our conclusions in Section 7. All results presented in this paper use a cosmology in which $\\Omega_{\\rm m, 0}$ = 0.3, $\\Omega_{\\rm \\Lambda, 0}$ = 0.7, and H$_0$ = 70 km s$^{-1}$ Mpc$^{-1}$. Errors quoted in this paper are 90 per cent confidence for one parameter of interest (i.e., $\\chi^2_{\\rm min}$ + 2.7), unless otherwise stated. ", "conclusions": "We have presented results from \\XMM observations of the nuclei of the radio galaxies 3C\\,305, DA\\,240, and 4C\\,73.08. We have shown the following: \\begin{enumerate} \\item The X-ray spectrum of the narrow-line FRII radio galaxy 4C\\,73.08 can be modeled as the sum of a heavily absorbed power law associated with a luminous accretion disk and circumnuclear obscuring structure, together with an unabsorbed component of X-ray emission that has a common origin with the radio emission at the base of an unresolved jet. This behavior is consistent with the other narrow-line FRII radio galaxies studied by \\cite{evans06} and \\cite{hec06}. \\item The nuclear X-ray spectrum of the FRII giant radio galaxy DA\\,240, optically classified as a low-excitation radio galaxy, can be modeled as a single, unabsorbed power law that is likely associated with emission from the parsec-scale jet. The upper limit to the X-ray luminosity of any additional, accretion-related emission suggests that the accretion process in DA\\,240 is substantially sub-Eddington and likely radiatively inefficient in nature. \\item The X-ray emission in the nucleus of the narrow-line radio galaxy 3C\\,305 can be modeled as an unabsorbed power law that originates at the base of the jet. However, it shows no evidence for heavily absorbed X-ray emission was found in the NLRGs studied by \\cite{evans06}. \\item We have discovered an X-ray counterpart to the NE hotspot of the giant radio galaxy DA\\,240. We argue that the emission process is overwhelmingly likely to be synchrotron emission. Because of the high X-ray flux of the hotspot, it is a good candidate for followup high-resolution X-ray observations. \\item We have discussed the different origins of optical emission lines in the nuclear and circumnuclear gaseous environments of radio galaxies. These include photoionization from the AGN accretion flow or parsec-scale jet, shock-excitation by the radio jet, or cooling gas in the centers of clusters. This may lead to occasional misclassification of genuinely weak-lined sources such as 3C\\,305 as high-excitation sources. \\item We therefore argue that there is not necessarily always a one-to-one correspondence between optical emission-line class (low- vs. high-excitation) and accretion-flow state (inefficient flow vs. standard thin disk), especially when low angular-resolution optical spectroscopy is used. We suggest that only the combination of high-resolution optical, X-ray, and infrared observations can reliably uncover the nature of the central engine in radio-loud AGN. \\end{enumerate}" }, "0808/0808.0349_arXiv.txt": { "abstract": "s{ A striking concentration of ultra-high energy cosmic ray (UHECR) events observed by the Pierre Auger Observatory around the direction of the nearby radio galaxy Centaurus A revives the idea that radio galaxies may be dominant sources of UHECR. In this paper, we give a brief overview about processes which may accelerate protons and nuclei in radio galaxies, and their relation to jet power, radio morphology and cosmic source density. We argue that, except for the most powerful FR-II radio galaxies, processes in radio lobes are unlikely to explain the origin of UHECR. However, Fermi acceleration of protons at internal shocks in the ``blazar-zone'' of \\textit{all} radio galaxies, and their photohadronic conversion into neutrons, may lead to the ejection of ``UHECR-beams'', which remain collimated over several Mpc. Consequences of this hypothesis for the interpretation of the UHECR event distribution, in particular for the special case of Centaurus A, are discussed.} ", "introduction": " ", "conclusions": "" }, "0808/0808.2385_arXiv.txt": { "abstract": "We present an analysis of six 12 ${\\umu}$m selected Seyfert 2 galaxies that have been reported to be unabsorbed in the X-ray. By comparing the luminosities of these galaxies in the mid-IR (12 ${\\umu}$m), optical ([O\\,{\\sc iii}]) and hard X-ray (2-10 keV), we show that they are all under-luminous in the 2-10 keV X-ray band. Four of the objects exhibit X-ray spectra indicative of a hard excess, consistent with a heavily obscured X-ray component and hence a hidden nucleus. In these objects the softer X-rays may be dominated by a strong soft scattered continuum or contamination from the host galaxy, which is responsible for the unabsorbed X-ray spectra observed, and accounts for the anomalously low 2-10 keV X-ray luminosity. We confirm this assertion in NGC 4501 with a {\\it Chandra} observation, which shows hard X-ray emission coincident with the nucleus, consistent with heavy absorption, and a number of contaminating softer sources which account for the bulk of the softer emission. We point out that such ``Compton thick\" sources need not necessarily present iron K$\\alpha$ emission of high equivalent width. An example in our sample is IRASF 01475-0740, which we know must host an obscured AGN as it hosts a hidden broad line region seen in scattered light \\citep{tran03}. The X-ray spectrum is nonetheless relatively unobscured and the iron K$\\alpha$ line only moderate in strength ($\\sim 160$ eV). These observations can be reconciled if the hidden nuclear emission is dominated by transmitted, rather than reflected X-rays, which can then be weak compared to the soft scattered light or galactic emission even at 6.4 keV. Despite these considerations, we conclude that two sources, NGC 3147 and NGC 3660, may intrinsically lack a broad line region (BLR), confirming the recent results of \\cite{bianchi08} in the case of NGC 3147. Neither X-ray spectrum shows signs of hidden hard emission and both sources exhibit X-ray variability leading us to believe we are viewing the nucleus directly. ", "introduction": "The current model of AGN unification (e.g. \\cite{antonucci93}) states that the absence of broad lines in the optical spectra of Seyfert 2 galaxies is due to their obscuration by an optically thick structure. Following this, the obscuration that blocks the BLR from view should also absorb soft X-rays from the continuum source in the same region as the BLR. This is well represented and documented in the prototypical Seyfert 2, NGC 1068, which shows an X-ray reflection spectrum and strong iron K$\\alpha$ emission \\citep{iwasawa97}, typical of Compton thick obscuration of the X-ray source, as well as displaying polarised broad lines (PBLs) in its optical spectrum, indicative of a hidden BLR \\citep{antonucci85}. The X-ray line of sight column density ($N_{\\rm{H}}$) has been measured to be in excess of $10^{23}$ cm$^{-2}$ in many Seyfert 2s \\citep{awaki91}, greatly exceeding the typical values in Seyfert 1s and lending strong support to the unification model. However, there remains an enigmatic group of Seyfert 2s that appear unabsorbed in X-rays (e.g. \\cite{pappa01}; \\cite{panessa02}), whose properties therefore seem to contradict the unified scheme. This apparent mismatch may result from a genuine absence of the BLR. This intriguing possibility has been proposed to pertain in Seyfert 2 galaxies with particularly low accretion rates, with concomitant implications for the origin of the broad optical lines \\citep{nicastro03}. Apparent confirmation of such a BLR-free AGN has recently been presented by Bianchi et al. (2008) who considered the case of NGC 3147. They were able to reject an alternative interpretation that the mismatch between X-ray and optical properties was due to variability. They also argued against a hidden nucleus based on the {\\it XMM-Newton} spectrum of NGC 3147, which shows a relatively modest iron K$\\alpha$ emission line. Most objects with very heavily obscured nuclei show very strong iron K$\\alpha$ lines \\citep{turner97} with the most extreme examples being the Compton thick sources, with lines often in excess of 1 keV equivalent width \\citep{matt96_2}. On the other hand, it is conceivable that in these unabsorbed sources we are simply being mislead by the X-ray data. The emission from the nucleus could be contaminated by diffuse or extra-nuclear point-source emission from the host galaxy or by AGN X-rays scattered by hot electrons. It is therefore possible that there are many Seyfert 2s which from their X-ray spectra, appear to be unabsorbed, but in fact are hiding a deeply buried AGN and that the unabsorbed profile is not necessarily nuclear. The observed equivalent width of the iron K$\\alpha$ line is strongly dependent on the relative contributions of the various components and, in addition, the geometry and physical parameters of the obscuring material. In this paper we present an analysis of six apparently unabsorbed Seyfert 2 galaxies, aiming to probe the nature of these objects. \\begin{table*} \\centering \\begin{minipage}{140mm} \\caption{Positional and observational data on the six Seyfert 2 galaxies featured. Spectropolarimetry data are taken from Tran (2003) and positional, redshift and distance data are taken from NED.}\\label{obsdat} \\begin{tabular}{|l l l l l l l c l l l} \\hline Object name \t& RA \t\t& Dec \t\t& z \t& Distance \t& $N_{\\rm{H}}$ \t\t& PBLs? & $\\frac{\\rm{[N\\,II]}}{\\rm{H\\alpha}}$ & $\\frac{\\rm{[O\\,III]}}{\\rm{H\\beta}}$ & Ref. \\\\ (1) \t\t& (2) \t\t& (3) \t\t& (4) \t& (5) \t\t& (6) \t\t\t& (7) \t& (8) \t& (9) \t& (10)\\\\ \\hline IRASF 01475-0740& 01h50m02.7s \t& -07d25m48s \t& 0.0177& 69.7 \t\t& $2.0 \\times 10^{20}$ \t& yes \t& 0.49 \t& 5.21 \t& 2 \\\\ NGC 3147 \t& 10h16m53.6s \t& +73d24m03s \t& 0.0094& 39.8 \t\t& $2.9 \\times 10^{20}$ \t& no \t& 2.71 \t& 6.11 \t& 1 \\\\ NGC 3486 \t& 11h00m23.9s \t& +28d58m29s \t& 0.0023& 13.5\t\t& $1.9 \\times 10^{20}$ \t& - \t& 1.05 \t& 4.57 \t& 1 \\\\ NGC 3660 \t& 11h23m32.3s \t& -08d39m31s \t& 0.0123& 56.2\t\t& $3.4 \\times 10^{20}$ \t& no \t& 0.82 \t& 2.63 \t& 3 \\\\ NGC 3976\t& 11h55m57.6s \t& +06d45m03s \t& 0.0083& 39.4\t\t& $1.1 \\times 10^{20}$ \t& - \t& 1.96 \t& 3.52 \t& 1 \\\\ NGC 4501\t& 12h31m59.2s \t& +14d25m14s \t& 0.0076& 36.0\t\t& $2.6 \\times 10^{20}$ \t& no \t& 2.10 \t& 5.25 \t& 1 \\\\ \\hline \\end{tabular} Col. (1) Galaxy name as given by \\cite{rush93} in the extended 12 micron galaxy sample; Col. (2) Right ascension (NED, J2000); Col. (3) Declination (NED, J2000); Col. (4) Redshift (NED); Col. (5) Luminosity distance (NED, Mpc); Col. (6) Galactic $N_{\\rm{H}}$ (NED, cm$^{-2}$); Col. (7) Indicates whether the objects have polarised broad lines in the optical spectrum if data exists \\citep{tran03}; Cols. (8-10) Line ratios as measured by (1) \\cite{ho97}; (2) \\cite{degrijp92}; (3) \\cite{gu06}. \\end{minipage} \\end{table*} \\section[]{Data Analysis} \\subsection[]{Sample Selection} The parent sample for our study is the extended {\\it IRAS} 12 $\\umu$m sample of \\cite{rush93}. From this we select objects defined as Seyfert 2s by the NASA/IPAC Extragalactic Database (NED). We perform our own post-hoc analysis of the optical spectra below. We require there to be good quality X-ray data ({\\it XMM-Newton, Chandra or ASCA}) with a column density indicating that they are unabsorbed in X-rays ($N_{\\rm{H}} < 10^{22}$ cm$^{-2}$), from the literature or our own analysis. The NED optical classifications may not necessarily be robust, so we searched the literature to deselect sources with incorrect or ambiguous optical classification. The remaining six objects, IRASF 01475-0740, NGC 3147, NGC 3486, NGC 3660, NGC 3976, NGC 4501 form our sample: they are all unambiguously classified as AGN using line ratio diagnostics (Line ratios and references given in Table \\ref{obsdat} and BPT diagram presented in Fig. \\ref{bptdiag}). \\subsection[]{Optical data} We compiled optical spectroscopy, spectropolarimetry and line ratio data for our objects from the literature and present these data with references in Table \\ref{obsdat} and Fig. \\ref{bptdiag}, which plots the ratio [O\\,{\\sc iii}] $\\lambda5007$/H$\\beta$ versus the ratio [N\\,{\\sc ii}] $\\lambda6584$/H$\\alpha$. We plot these on top of the catalogue of low redshift SDSS galaxies of \\cite{kauffmann03}. The demarcation line is that defined by \\cite{kewley01} separating AGN and starbursts, hence showing that their AGN classification is not in doubt. Additionally, we plot the line ratios of NGC 6810, a Seyfert 2 also unabsorbed in X-rays, but shown to have a dubious Seyfert 2 classification due to broader-than-normal optical lines produced by a super-wind \\citep{strickland07}. The authors used an $XMM-Newton$ observation to show that the X-ray emission from this galaxy was probably due to X-ray binaries. \\begin{figure} \\includegraphics[width=67mm,angle=90]{bptdiag.ps} \\caption{BPT diagram (Baldwin et al. 1981) of the six featured Seyfert 2 (green diamonds) sources plus NGC 6810 (red triangle), which has previously but incorrectly been classified as a Seyfert 2 (Strickland 2007), plotted on top of 55, 757 low redshift SDSS galaxies (blue dots) of Kauffmann et al. (2003). The demarcation line is that defined by Kewley et al. (2001) and separates AGN from starbursts.}\\label{bptdiag} \\end{figure} \\subsection[]{X-ray Data} We carried out the X-ray analysis on {\\it XMM-Newton} observations of IRASF 01470-0740, NGC 3147, NGC 3486, NGC 3976 and NGC 4501; {\\it Chandra} observations of NGC 3147 and NGC 4501 and an {\\it ASCA} observation of NGC 3660 (X-ray observational information listed in table \\ref{xobsdat}). For {\\it XMM-Newton} data we use {\\sc sas v7.0} tasks to perform spectral extractions on the EPIC-pn data, for the {\\it Chandra} ACIS-S data we use {\\sc ciao v3.4} tasks and use {\\it ASCA} data products from the {\\it Tartarus} database. All spectra were grouped with a minimum of 20 counts per bin, with the exception of the {\\it Chandra} observation of NGC 4501, where the spectrum was grouped using a minimum of 7 counts. Spectral fitting was carried out using {\\sc xspec v11.3}. \\begin{table} \\centering \\caption{Information on the X-ray observations used in this analysis. Exposure time in brackets give the time after filtering for background flares in {\\it XMM-Newton} data.}\\label{xobsdat} \\begin{tabular}{l l l l l} \\hline Object name\t& Observatory\t\t& Date\t\t& Obs. ID\t\t& exp.\t\t\\\\ &\t\t\t&\t\t&\t\t\t& (ks)\t\t\\\\ \\hline F01475-0740 \t& {\\it XMM-Newton}\t& 2004-02-01\t& 0200431101\t\t& 12(9) \t\\\\ NGC 3147 \t& {\\it XMM-Newton}\t& 2006-10-06\t& 0405020601\t\t& 17(14)\t\\\\ & {\\it Chandra}\t\t& 2001-09-19\t& 1615\t\t\t& 2\t\t\\\\ NGC 3486 \t& {\\it XMM-Newton} \t& 2001-05-09\t& 0112550101\t\t& 15(4) \t\\\\ NGC 3660 \t& {\\it ASCA}\t\t& 1995-06-09\t& 73039000\t\t& 26 \t\t\\\\ NGC 3976\t& {\\it XMM-Newton}\t& 2006-06-16\t& 0301651801\t\t& 14(4) \t\\\\ NGC 4501\t& {\\it XMM-Newton} \t& 2001-12-04\t& 0112550801\t\t& 14(3) \t\\\\ & {\\it Chandra}\t\t& 2002-12-09\t& 2922 \t\t\t& 18 \t\t\\\\ \\hline \\end{tabular} \\end{table} \\section[]{Estimates of Bolometric Luminosity} We estimate the bolometric luminosities, $L_{\\rm{Bol}}$, of the six AGN from their 12 ${\\umu}$m and [O\\,{\\sc iii}] fluxes using bolometric corrections, $\\kappa_{\\rm 12\\umu m}$ calculated from template SEDs of \\cite{mrr08} and $\\kappa_{\\rm [O\\,III]}$ published by \\cite{heckman04}. We also estimate bolometric luminosities from the unabsorbed 2-10 keV (HX) luminosity using the mean bolometric correction, $\\kappa_{\\rm HX}$, of \\cite{vasudevan07}. Table \\ref{lumins} presents the observed luminosities with these estimates and Fig. \\ref{lumfig} plots the estimated bolometric luminosities from the [O\\,{\\sc iii}] luminosity against the estimated bolometric luminosities from the 2-10 keV luminosity. From this analysis, all six objects appear to be significantly under-luminous in the 2-10 keV X-ray band by factors of 10-100. For a typical Seyfert 2 this is easily understood, as we expect the 2-10 keV X-rays to be suppressed by absorption, but of course for our sample it seems the evidence for that absorption in terms of the spectral shape is absent. As discussed below, despite the lack of any obvious soft X-ray absorption in these objects, some of the X-ray spectra present evidence for hidden hard components. This allows an additional estimate of the intrinsic AGN bolometric luminosity show as the red arrows in Fig. \\ref{lumfig}. These are discussed in more detail below. \\begin{table*} \\centering \\begin{minipage}{140mm} \\caption{Observed luminosities of the sample using distances from Table \\ref{obsdat}. $12 \\umu$m luminosity is calculated from the {\\it IRAS} flux density presented by Rush et al. 1993 and [O\\,{\\sc iii}] luminosities have been corrected for absorption using the formulation of Bassani et al. (1999). The respective predicted bolometric luminosity calculated from these measured luminosities are also given.}\\label{lumins} \\begin{tabular}{|l | c c c|c c c|} \\hline Object name & \\multicolumn{3}{|c|}{Observed luminosities (erg s$^{-1}$)} & \\multicolumn{3}{|c|}{$L_{\\rm{Bol}}$ (erg s$^{-1}$) predicted from:}\\\\ &$L_{\\rm{12{\\umu}m}}$ & $L_{\\rm [O\\,III]}$ & $L_{\\rm{HX}}$ & $L_{\\rm{12{\\umu}m}}$ & $L_{\\rm [O\\,III]}$ & $L_{\\rm{HX}}$\\\\ \\hline IRASF 01475-0740& $7.68\\times 10^{43}$ & $4.80\\times 10^{41}$ & $4.71\\times 10^{41}$ & $8.54\\times 10^{44}$ & $1.68\\times 10^{45}$ & $1.28\\times 10^{43}$\\\\ NGC 3147 \t& $5.02\\times 10^{43}$ & $1.50\\times 10^{40}$ & $2.68\\times 10^{41}$ & $5.58\\times 10^{44}$ & $5.08\\times 10^{43}$ & $7.25\\times 10^{42}$\\\\ NGC 3486 \t& $2.83\\times 10^{42}$ & $3.89\\times 10^{38}$ & $3.26\\times 10^{39}$ & $3.14\\times 10^{43}$ & $1.36\\times 10^{42}$ & $8.82\\times 10^{40}$\\\\ NGC 3660 \t& $3.15\\times 10^{43}$ & $9.38\\times 10^{40}$ & $8.72\\times 10^{41}$ & $3.50\\times 10^{44}$ & $3.28\\times 10^{44}$ & $2.36\\times 10^{43}$\\\\ NGC 3976 \t& $1.59\\times 10^{43}$ & $4.30\\times 10^{39}$ & $2.18\\times 10^{40}$ & $1.76\\times 10^{44}$ & $1.50\\times 10^{43}$ & $5.90\\times 10^{41}$\\\\ NGC 4501 \t& $6.79\\times 10^{43}$ & $9.57\\times 10^{39}$ & $1.26\\times 10^{40}$ & $7.54\\times 10^{44}$ & $3.35\\times 10^{43}$ & $3.42\\times 10^{41}$\\\\ \\hline \\end{tabular} \\end{minipage} \\end{table*} \\begin{figure} \\includegraphics[width=67mm,angle=90]{lumplot.ps} \\caption{Estimates of the bolometric luminosities from $L_{{\\rm [O\\,III]}}$ with $1\\sigma$ variance (green lines) against the estimated bolometric luminosities from $L_{\\rm{HX}}$. The solid black line displays where on the plot these estimates would lie should they be equal, as expected, and the dashed black lines represent the $1\\sigma$ spread in $\\kappa_{\\rm HX}$. As these objects are all positioned under the solid line, they are all under-luminous in X-rays. Red arrows show how the bolometric luminosity predicted from $L_{\\rm{HX}}$ changes when we try to estimate the intrinsic $L_{\\rm{HX}}$ of the nucleus based on the X-ray spectral fitting.}\\label{lumfig} \\end{figure} \\section[]{X-Ray Analysis} \\subsection{IRASF 01475-0740} The {\\it XMM-Newton} spectrum of IRASF 01475-0740 is well fitted by a power-law absorbed by a column of $N_{\\rm{H}} = 4.1 \\times 10^{21}$ cm$^{-2}$. While this will produce some reddening and extinction in the optical, it is unlikely to be sufficient to suppress the optical broad lines entirely unless the absorber has an anomalously high dust-to-gas ratio (c.f. MCG-6-30-15; \\cite{reynolds97}). The X-ray spectrum also shows line emission between 6 and 7 keV. The centroid energy of this line is not well constrained, so we assume that the emission is from neutral iron K$\\alpha$ at a rest energy of 6.4 keV, though it may originate from ionised iron. A gaussian fit to this feature with the centroid energy fixed at 6.4 keV gives an equivalent width (EW) of 160 eV. On the face of it, this relatively modest iron K$\\alpha$ EW argues against a hidden Compton thick nucleus (Bianchi et al. 2008). We found, however, that it was possible to fit a second, heavily absorbed component from new Monte-Carlo models of X-rays in heavily obscured AGN, incorporating line emission (Brightman et al., in preparation), similar to those presented by e.g. \\cite{ghisellini94} and \\cite{krolik94}. For simplicity given the limitations of the data we fit the extreme case of 4$\\pi$ coverage (i.e. a spherical distribution of matter), and we constrained the column density of the heavily absorbed component to $N_{\\rm{H}}=2 \\times 10^{24}|^{+65}_{-0.8}$ cm$^{-2}$. Although it is not formally required in the fit, this demonstrates a clear scenario in which the multiwavelength data can be reconciled, as such a column is easily sufficient to suppress the direct nuclear emission in the optical. By correcting for absorption in this heavily obscured component, we can make another prediction of the intrinsic luminosity of the nucleus and calculate the corresponding bolometric luminosity. This now agrees well with estimations from $L_{\\rm [O\\,III]}$ and $L_{\\rm{12{\\umu}m}}$. The {\\it XMM-Newton} observation also shows no clear signs of variability. The spectropolarimetric survey of \\cite{tran03} reveals that this AGN hosts a HBLR, in full agreement with our conclusion of a heavily buried nuclear contribution. The same electrons which scatter the broad optical lines into the line of sight can then also scatter nuclear X-rays accounting for the apparently unobscured nature of the soft spectrum. \\subsection{NGC 3147} Our analysis of the {\\it XMM-Newton} spectrum of NGC 3147 shows no absorption above the Galactic column, however it features a small (130 eV) iron K$\\alpha$ line at 6.4 keV. These results are in full agreement with those of Bianchi et al. (2008). \\cite{bianchi08} used the small EW of the iron K$\\alpha$ line to argue against the Compton thick nature of this source and came to the conclusion that NGC 3147 has an intrinsic absence of a BLR. However, as we have shown in IRASF01475-0740, this conclusion is not necessarily robust. Fitting the Monte Carlo models to the spectrum once again shows that a very heavily obscured component is consistent with the spectrum. Here the column density is constrained to be $N_{\\rm{H}}=9 \\times 10^{23}|^{+419}_{-1.0}$ cm$^{-2}$, with the main constraint coming from the iron K$\\alpha$ line. We can again use this second component to predict the intrinsic luminosity of this source and calculate the corresponding bolometric luminosity, which also agrees well with estimations from $L_{\\rm [O\\,III]}$ and $L_{\\rm{12{\\umu}m}}$. Other evidence suggests, however, that this may not be the correct interpretation in this case. The 2ks {\\it Chandra} observation of NGC 3147 shows that the measured flux of the nucleus is $\\sim2$ times that measured by {\\it XMM-Newton} indicating that it is variable in nature and hence it must be being observed directly, rather than in scattered light. Furthermore, the {\\it Chandra} imaging reveals no evidence for extra-nuclear X-ray sources that could be producing the unabsorbed X-ray profile or the variability, and spectral extraction from the nuclear region in the {\\it Chandra} image does not reveal a significantly harder X-ray spectrum then the larger {\\it XMM-Newton} beam (Table~4). \\subsection{NGC 3486} The {\\it XMM-Newton} spectrum of NGC 3486 is well fitted by a power-law absorbed by the Galactic column only and so appears to be another unabsorbed Seyfert 2 galaxy. However, an excess at hard energies points to a different scenario. The spectrum is also well fitted by the addition of a Compton reflection model ({\\tt pexmon}, \\cite{nandra07}) with an underlying thermal component ({\\tt raymond}), so this could in fact be a Compton thick Seyfert 2 which looks unabsorbed. We also attempt to add a strongly absorbed transmission component to the model, which also improves the fit from a simple power-law. However, the column density of this component is poorly constrained, so we fix this to an arbitrary log($N_{\\rm{H}}$) = 24.5. Estimates of the intrinsic $L_{\\rm{HX}}$ from the reflection component predict a $L_{\\rm{Bol}}$ which agrees well with $L_{\\rm{Bol}}$ estimated from $L_{\\rm [O\\,III]}$, suggesting that NGC 3486 is under-luminous in X-rays due to Compton thick obscuration. Finally, there is no evidence for variability of the X-ray flux in the {\\it XMM-Newton} observation. \\subsection{NGC 3660} As there have been no observations of NGC 3660 with {\\it XMM-Newton} or {\\it Chandra}, we use the {\\it ASCA} observation and its documentation in the {\\it Tartarus} catalogue. The {\\it ASCA} spectrum is fitted well by a power-law with no absorption above the Galactic column. Fig. \\ref{ngc3660lc} shows the light-curve for the 26 ks observation which shows significant variability on short time-scales. \\begin{figure} \\centering \\includegraphics[width=84mm]{ngc3660_lc.ps} \\caption{{\\it ASCA} lightcurve of NGC 3660 showing the X-ray variable nature of the source.}\\label{ngc3660lc} \\end{figure} \\subsection{NGC 3976} The {\\it XMM-Newton} spectrum of NGC 3976 is very similar to that of NGC 3486 as it shows no intrinsic absorption, but a hard excess present above the simple power-law fit allows us to add a heavily absorbed power-law component. A strongly absorbed transmission component produces an overall better fit, suggesting that NGC 3976 is also hiding a heavily obscured nucleus beneath its apparently unabsorbed soft X-ray spectrum. Again, there is also no variability detected in this observation. \\subsection{NGC 4501} The {\\it XMM-Newton} spectrum of NGC 4501 reveals another apparently unabsorbed Seyfert 2 galaxy in X-rays. It is well fitted by a power-law with no intrinsic absorption plus emission from a thermal plasma component, but shows no clear hard excess or other spectral evidence supporting a deeply buried AGN (such as iron K$\\alpha$ emission). An entirely different picture emerges, however, when one considers the {\\it Chandra} data, which have higher spatial resolution. The {\\it Chandra} image of NGC 4501 reveals a hard X-ray emission coincident with the optically defined nucleus (Fig. \\ref{chanimg}) consistent with heavy X-ray absorption. It also shows that there are also multiple extra-nuclear sources present, including diffuse soft emission close to the nucleus, which {\\it XMM-Newton} could not resolve. The implication is that the unabsorbed nature of the {\\it XMM-Newton} spectrum, which has a larger beam size, is due to contamination, and that the true nuclear emission is heavily obscured. We performed a spectral extraction of the optically defined nucleus using the region identified in Fig \\ref{chanimg} which did indeed reveal a hard excess above the unabsorbed power-law, as in NGC 3486 and NGC 3976, which we fit with a Compton reflection component ({\\tt pexmon}). Using the reflection component to estimate the intrinsic $L_{\\rm{HX}}$ shows that the X-ray faintness of NGC 4501 is probably due to heavy absorption (Fig. \\ref{lumfig}). There is also no evidence for variability of NGC 4501, either between subsequent observations by {\\it XMM-Newton} and {\\it Chandra}, or during them. \\begin{figure} \\centering \\includegraphics[width=50mm]{NGC4501_img.ps} \\caption{Chandra image of NGC 4501 smoothed by a $\\sigma=3$ pixels Gaussian. Red represents 0.5-1.0 keV emission, green represents 1.0-3.0 keV emission and blue represents 3.0 - 10.0 keV in a log scale. The black cross marks the position of the optical nucleus, the smaller green circle marks the {\\it Chandra} extraction region and larger green circle marks the equivalent {\\it XMM-Newton} extraction region. The white area coincident with the nucleus indicates that it is the source of hard X-ray emission, and therefore likely to be heavily absorbed.}\\label{chanimg} \\end{figure} \\begin{table*} \\centering \\begin{minipage}{140mm} \\caption{X-ray spectral fitting data for the sample, where R is the ratio of the underlying transmitted or reflected component to the power-law}\\label{fitdat} \\begin{tabular}{l l l l l l l l l l r@{}l} \\hline Object name \t& Model\t& {$N_{\\rm{H,1}}$} \t& {$N_{\\rm{H,2}}$} \t& $\\Gamma$ \t\t& EW$_{6.4}$ \t& kT \t\t\t& R\t\t\t& $\\chi_r^2/\\nu$& {$F_{\\rm HX}$}& \\multicolumn{2}{c}{$L_{\\rm int}$}\\\\ (1)\t\t& (2)\t& (3) \t\t\t& (4)\t \t\t& (5) \t\t\t& (6) \t\t& (7)\t\t\t& (8)\t\t\t& (9)\t\t& (10) \t\t& & (11)\t\\\\ \\hline IRASF01475-0740 & A\t& $0.41^{+0.07}_{-0.05}$& -\t\t\t& $2.02^{+0.06}_{-0.06}$& $160^{+300}_{-160}$& -\t\t& -\t\t\t& 65.5/58\t& 0.84\t\t& 49 &\t\t\\\\ ({\\it XMM-Newton})& C\t& $0.41^{+2.14}_{-0.18}$& -\t\t\t& $2.15^{+0.45}_{-0.24}$& - \t\t& - \t\t\t& $1.0^{+32}_{-1.0}$\t& 64.4/56 \t& 0.84\t\t& 48 &\t\t\\\\ & D\t& $0.46^{+0.17}_{-0.17}$& $200^{+6500}_{-80}$\t& $2.16^{+0.26}_{-0.23}$& - \t\t& - \t\t\t& $10^{+1600}_{-10}$\t& 63.3/56\t& 0.82\t\t& 465 &\t\t\\\\ \\hline NGC 3147 \t& A\t& $0.03^{+0.01}_{-0.01}$& -\t\t\t& $1.60^{+0.06}_{-0.06}$& $130^{+80}_{-100}$& -\t\t\t& -\t\t\t& 103/113\t& 1.4 \t\t& 27 &\t\t\\\\ ({\\it XMM-Newton})& C\t& $0.03^{+0.04}_{-0.01}$& -\t\t\t& $1.63^{+0.09}_{-0.07}$& - \t\t& - \t\t\t& $1.6^{+8.3}_{-1.6}$\t& 103/112\t& 1.4 \t\t& 42 &\t\t\\\\ \\vspace{0.1in}\t& D\t& $0.04^{+0.01}_{-0.02}$& $90^{+4210}_{-10}$\t& $1.68^{+0.04}_{-0.11}$& -\t\t& - \t\t\t& $1.9^{+10}_{-1.9}$\t& 102/112\t& 1.4 \t\t& 49 &\t\t\\\\ NGC 3147\t& A\t& $<0.07$\t\t& -\t\t\t& $1.50^{+0.19}_{-0.12}$& -\t\t& -\t\t\t& -\t\t\t& 62.2/49\t& 3.2\t\t& 61 &\t\t\\\\ ({\\it Chandra})\t&\t&\t\t\t& \t\t\t&\t\t\t&\t\t&\t\t\t&\t\t\t&\t\t&\t\t& &\t\t\\\\ \\hline NGC 3486 \t& A\t& $0.03^{+0.17}_{-0.03}$& -\t\t\t& $2.44^{+1.17}_{-0.61}$& - \t\t& - \t\t\t& -\t\t\t& 28.3/18\t& 0.04\t\t& 0 & .10\t\\\\ ({\\it XMM-Newton})& B\t& $<0.10$\t\t& -\t\t\t& $1.73^{+1.29}_{-0.83}$& -\t\t& $0.23^{+0.13}_{-0.09}$& -\t\t\t& 24.8/16\t& 0.09\t\t& 0 & .20\t\\\\ & C\t& $<0.06$\t\t& -\t\t\t& $1.9^{\\dag}$\t\t& -\t\t& $0.24^{+0.19}_{-0.12}$& $34^{+220}_{-34}$\t& 21.5/15\t& 0.15\t\t& 4 & .8\t\\\\ & D\t& $<0.06$\t\t& 320$^{\\dag}$\t\t& $1.9^{\\dag}$\t\t& -\t\t& $0.24^{+0.66}_{-0.10}$& $200^{+650}_{-200}$\t& 23.2/16\t& 0.09\t\t& 31 & \t\t\\\\ \\hline NGC 3660 \t& A\t& $0.03^{\\dag}$\t\t& -\t\t\t& $1.82^{+0.04}_{-0.04}$& - \t\t& - \t\t\t& -\t\t\t& 301/304\t& 2.3\t\t& 87 &\t\t\\\\ ({\\it ASCA})\t&\t&\t\t\t& \t\t\t&\t\t\t&\t\t&\t\t\t&\t\t\t&\t\t&\t\t& &\t\t\\\\ \\hline NGC 3976\t& A\t& $0.06^{+0.15}_{-0.06}$& -\t\t\t& $1.80^{+1.15}_{-0.60}$& - \t\t& - \t\t\t& -\t\t\t& 38.0/33\t& 0.07\t\t& 1 & .3\t\\\\ ({\\it XMM-Newton})& B\t& $0.10^{+0.10}_{-0.07}$& -\t\t\t& $1.9^{\\dag}$\t\t& -\t\t& $0.44^{+0.29}_{-0.18}$& -\t\t\t& 26.2/32\t& 0.05\t\t& 0 & .97\t\\\\ & C \t& $0.09^{+0.09}_{-0.06}$& -\t\t\t& $1.9^{\\dag}$\t\t& - \t\t& $0.43^{+0.83}_{-0.14}$& $43^{+49}_{-43}$\t& 27.0/30 \t& 0.10\t\t& 37 &\t\t\\\\ & D\t& $0.10^{+0.06}_{-0.04}$& $17^{+190}_{-12}$\t& $1.9^{\\dag}$\t\t& -\t\t& $0.44^{+0.18}_{-0.11}$& $3.0^{+25}_{-3.0}$\t& 24.8/33\t& 0.12\t\t& 4 & .2\t\\\\ \\hline NGC 4501\t& A \t& $0.21^{+0.33}_{-0.14}$& -\t\t\t& $3.91^{+2.76}_{-1.14}$& - \t\t& - \t\t\t& -\t\t\t& 21.2/13\t& 0.02\t\t& 0 & .36\t\\\\ \\vspace{0.1in} ({\\it XMM-Newton})& B\t& $0.02^{+0.10}_{-0.02}$& -\t\t\t& $2.13^{+1.14}_{-0.63}$& -\t\t& $0.36^{+0.20}_{-0.08}$& -\t\t\t& 6.82/11\t& 0.08\t\t& 1 & .2\t\\\\ NGC 4501\t& A\t& $<0.02$\t\t& -\t\t\t& $1.49^{+0.29}_{-0.36}$& -\t\t& -\t\t\t& -\t\t\t& $44.7^C$\t& 0.05\t\t& 0 & .78\t\\\\ ({\\it Chandra})\t& B\t& $1.75^{+0.12}_{-0.31}$& -\t\t\t& $1.65^{+0.25}_{-0.31}$& -\t\t& $0.05^{+0.01}_{-0.01}$& -\t\t\t& $22.7^C$\t& 0.07\t\t& 1 & .1\t\\\\ & C \t& $0.66^{+1.11}_{-0.66}$& -\t\t\t& $1.9^{\\dag}$\t\t& -\t\t& $0.39^{+0.16}_{-0.10}$& $29^{+32}_{-20}$\t& $8.64^C$\t& 0.08\t\t& 13 &\t\t\\\\ & D \t& $0.16^{+0.47}_{-0.16}$& $18^{+47}_{-15}$\t& $1.9^{\\dag}$\t\t& -\t\t& $0.35^{+0.34}_{-0.17}$& $8.9^{+28}_{-6.4}$\t& $10.7^C$\t& 0.09\t\t& 2 & .6\t\\\\ \\hline \\end{tabular} Notes: HX = 2-10 keV; $^\\dag$ indicates a fixed parameter; $^C$ indicates use of Cash statistics in spectral fitting; Col. (1) Galaxy name as given by \\cite{rush93}; Col. (2) Model used for fitting, A=pl, B=pl+therm, C=pl+therm+refl, D=pl+therm+trans; where pl = simple power-law; trans = Monte-Carlo model of transmitted component; therm = thermal component (raymond); refl = pure reflection component (pexmon); Col. (3) Column density of simple power-law, units of $10^{22}$ cm$^{-2}$; Col. (4) Column density of transmitted component, units of $10^{22}$ cm$^{-2}$; Col. (5) Power-law index, pegged for simple power-law and transmitted component; Col. (6) Equivalent width of the neutral iron line at 6.4 keV if present in eV; Col. (7) Temperature of the thermal component in keV; Col. (8) The normalisation ratio of the reflection or transmitted components to the simple power-law; Col. (9) Reduced chi-squared and number of degrees of freedom; Col. (10) Absorption corrected observed flux in the hard (2-10 keV) band, units of $10^{-12}$ erg cm$^{-2}$ s$^{-1}$; Col. (11) Hard X-ray intrinsic luminosity, units of $10^{40}$ erg s$^{-1}$, calculated from reflection or transmitted component where present. \\end{minipage} \\end{table*} \\begin{figure*} \\begin{minipage}{160mm} \\includegraphics[width=80mm]{IRASF01475_mo.ps} \\includegraphics[width=80mm]{NGC3147_mo.ps} \\includegraphics[width=80mm]{IRASF01475_data.ps} \\includegraphics[width=80mm]{NGC3147_data.ps} \\includegraphics[width=80mm]{NGC3147_chan_mo.ps} \\includegraphics[width=80mm]{NGC3486_mo.ps} \\includegraphics[width=80mm]{NGC3147_chan_data.ps} \\includegraphics[width=80mm]{NGC3486_data.ps} \\end{minipage} \\end{figure*} \\begin{figure*} \\begin{minipage}{160mm} \\includegraphics[width=80mm]{NGC3660_mo.ps} \\includegraphics[width=80mm]{NGC3976_mo.ps} \\includegraphics[width=80mm]{NGC3660_data.ps} \\includegraphics[width=80mm]{NGC3976_data.ps} \\includegraphics[width=80mm]{NGC4501_mo.ps} \\includegraphics[width=80mm]{NGC4501_chan_mo.ps} \\includegraphics[width=80mm]{NGC4501_data.ps} \\includegraphics[width=80mm]{NGC4501_chan_data.ps} \\caption{{\\it XMM-Newton} (unless otherwise stated) spectra of IRASF 01475-0740, NGC 3147, NGC 3486, NGC 3660, NGC 3976 and NGC 4501, shown with best fit model from Table \\ref{fitdat}. In the plots of the best fit model, blue dashed lines represent the unabsorbed power-law, red dot-dashed lines represent the thermal component if present, and the magenta dotted lines represent the reflection or transmitted component if present.} \\label{specfig} \\end{minipage} \\end{figure*} ", "conclusions": "\\subsection{Main Findings} We have presented X-ray and multi-waveband data for a sample of bona fide Seyfert 2 galaxies, as defined by their optical line ratios, which appear unabsorbed in the X-ray. Our first key finding is that the 2-10 keV X-ray luminosities of the objects are low with respect to their mid-IR and optical [O\\,{\\sc iii}] luminosities. Under normal circumstances this would be interpreted straightforwardly as being due to suppression of the X-ray flux by an absorbing column, however a naive analysis of the X-ray spectra is at odds with this interpretation. It may be that the simplest interpretation of the spectra is misleading us: if scattered or host galaxy emission dominates the spectrum below $\\sim 5$ keV, the spectrum may appear unabsorbed when in fact a harder nuclear components is present. We favour the latter interpretation in four objects. These show tentative evidence for a hard excess which is likely to be a hidden reflection or transmission spectrum indicative of a heavily obscured nucleus, accounting for the low hard X-ray luminosity of these objects. The most compelling case is NGC 4501, where the high resolution {\\it Chandra} image reveals a hard source co-incident with the nucleus embedded in a number of softer emitting regions. The clear conclusion in this case is that is that 2-10 keV {\\it XMM-Newton} data are dominated by non-nuclear emission, accounting for the unabsorbed X-ray spectrum. We infer a similar conclusion for IRAS F01475-0740, NGC 3486 and NGC 3976. For two of our sources (NGC 3147 and NGC 3660) we see no signs of hidden reflection or transmission in the {\\it XMM-Newton} X-ray spectra. NGC 3147 shows a single soft X-ray source co-incident with the nucleus in the Chandra image, and both exhibit X-ray variability. These observations clearly point to the idea that we are seeing the nuclei directly in these sources, so the evidence suggest that they might intrinsically lack a BLR. It therefore appears from our analysis that there are two populations of unabsorbed Seyfert 2 galaxies and that they appear unabsorbed in X-rays for different reasons entirely. It may be that our view of the AGN is genuinely unobscured, and that the lack of broad lines in their optical spectra is due to an intrinsic absence or weakness of the BLR. Alternatively though, it may be that the X-ray spectrum is misleading, and that the unabsorbed spectra are due to scattered or galactic soft X-rays which dominate the emission from a much more powerful obscured nucleus. We discuss each in turn. \\subsection{Genuinely unobscured Seyfert 2s} \\subsubsection{NGC 3147} NGC 3147 has no hidden broad line region and a {\\it Chandra} observations shows no discernible extra-nuclear sources. Between observations by {\\it Chandra} and {\\it XMM-Newton}, the hard X-ray flux drops by a factor of $\\sim 2$ indicating that it is variable in X-rays which means that we are likely to be seeing the nucleus of NGC 3147 directly. \\cite{bianchi08} carried out simultaneous optical/X-ray observations of NGC 3147 to show that the apparent mismatch between optical and X-ray classification was not due to differential variability. By noting a small equivalent width of the iron line and a large ratio between hard X-ray and [O {\\sc iii}] fluxes, they come to the same conclusion that the nucleus of NGC 3147 is genuinely unobscured, and that this AGN must therefore intrinsically lack a broad line region. We cannot however rule out the Compton thick nature of this source without observing it above 10 keV, where transmission from a Compton thick medium would dominate. \\subsubsection{NGC 3660} The X-ray variability of this source on short time-scales and unabsorbed profile of its X-ray spectrum leads us to the conclusion that we are genuinely viewing the nucleus of NGC 3660 directly. This leads us to believe that NGC 3660 also lacks a broad line region as we believe with NGC 3147. However, the origin of the X-ray variability is not necessarily nuclear as variability was recently discovered in the ultra-luminous X-ray (ULX) source, M82 X-1 \\citep{mucciarelli06}. A {\\it Chandra} observation of NGC 3660 should be able to resolve any ULXs as we have shown for NGC 4501. We also cannot rule out differential variability as the cause of the mismatch, as \\cite{bianchi08} did for NGC 3147. As has been observed in other X-ray variable Seyferts (e.g. NGC 4151, \\cite{gaskell86}), the broad line flux has also been seen to vary. If the broad line flux is observed in its low state, the object will be seen as type 2, despite being unobscured. NGC 3660 will also need to be observed simultaneously in the optical and in the X-rays to rule out this possibility. \\subsection{Misidentified Heavily Obscured Seyfert 2s} \\subsubsection{IRASF 01475-0740} \\cite{bianchi08} use the small size of the iron line in NGC 3147 to argue against the Compton thick nature of that source. However, IRASF01475-0740 also shows a small iron line, but in this object we know that heavy nuclear obscuration is occurring as shown by \\cite{tran03}. This spectropolarimetric study showed that IRASF01475-0740 has broad lines in its polarised optical spectrum, which are missing in its normal optical spectrum, confirming that our line of sight to the nucleus is blocked by optically thick material. We show that a Compton thick source may exhibit only a moderate iron line if the continuum emission below 10 keV is dominated by a strong scattered continuum or extra-nuclear emission. We demonstrated this by adding a second component to the fit of the X-ray spectrum, with a much larger column density than the first component, by using the iron line as a constraint. The intrinsic hard X-ray luminosity of this second component is a factor of $\\sim 10$ greater than the observed luminosity. Assuming that the observed hard X-ray spectrum is indeed dominated by scattered light, we calculate an upper limit on the scattering fraction by fixing the column density of the second component to its lower limit. The upper limit on the scattering fraction turns out to be $50\\%$, far higher than the typical $\\sim 3\\%$ seen in Seyfert 2 galaxies \\citep{cappi06}. \\cite{ueda07} observed a $<1\\%$ scattered fraction in two Compton thick AGN, SWIFT J0601.98636 and SWIFT J0138.64001. They used that fraction to suggest that these objects have a large covering fraction of the torus, or a low abundance of the gas responsible for the scattering in those objects. The inverse could be true for IRASF0175-0740 - it may have a low covering fraction of the torus, or a high abundance of the gas responsible for the scattering. Scattered components are also often seen in reflection dominated sources (eg. NGC 1068; \\citep{pier94}). If there is an underlying reflection component in IRASF0175-0740, this may also be dominated by scattered emission if the inclination angle of the torus is very high so that only a small part of the inner wall of the torus is directly visible and not self obscured. This would give a very high scattered/reflected fraction as noted for model C in Table \\ref{fitdat}. If the line emission seen here between 6 and 7 keV in fact originates from ionised iron, then we could be seeing reflection from ionised matter, as seen by \\cite{nandra07_2} in IRAS 00182-7112. The effect would be to increase the soft X-ray flux compared to neutral reflection as the lighter elements responsible for absorbing the soft X-rays will have been stripped of most of their electrons. This would fit well with the soft spectrum seen in IRASF01475-0740. The case of IRASF01745-0740 shows that Compton thick/intermediate AGN do not have to be reflection dominated with intense iron lines as they can be transmission dominated with a strong scattered component and no intense iron lines. \\subsubsection{NGC 4501} By constraining hard X-ray emission to the nucleus in the {\\it Chandra} observation of NGC 4501 we confirmed that this source is in fact Compton thick. Since the extra-nuclear sources seen in the {\\it Chandra} image lie outside the {\\it XMM-Newton} spectral extraction region used, we surmise that the power-law and thermal emission which dominate over the reflection component in the {\\it XMM-Newton} spectrum does not originate from ULXs or the AGN, but from diffuse galactic emission close to the nucleus. If this is scattered emission from the nucleus, the scattered fraction is $\\sim 3\\%$ in this Seyfert. \\subsubsection{NGC 3486} Fitting the hard excess in the X-ray spectrum of NGC 3486 with a reflection model yields an improved fit to this data, but the significance is too low to conclude that the source is Compton thick, as a reflection spectrum would imply. However, the {\\it XMM-Newton} spectrum here, is in all ways similar to the {\\it Chandra} spectrum of NGC 4501, which we showed to originate from a hard nucleus which is very likely Compton thick. As NGC 3486 is severely under-luminous in hard X-rays when compared to optical and mid-infrared luminosities, this conclusion seems likely for NGC 3486 too. Again we need to observe this Seyfert 2 above 10 keV to confirm our hypothesis, or with the high angular resolution capabilities of {\\it Chandra}. \\subsubsection{NGC 3976} As with NGC 3486, the hard excess present in this X-ray spectrum is well fitted by the addition of a reflection component. However, it is also well fitted by a transmission component. Again, this spectrum is very similar to the {\\it Chandra} spectrum of NGC 4501, so we conclude that this source is also heavily obscured. \\subsection{Wider Implications} The discovery of AGN which do not have a broad line region has important consequences for the AGN unification model. We can no longer exclusively invoke orientation based effects to explain the difference between type 1 and type 2 AGN. There must be some other physical effect which means that the BLR is absent. \\cite{nicastro03} studied a sample of Seyfert 2s extracted from the spectropolarimetric study of \\cite{tran03} and concluded that the presence of a hidden broad line region in Seyfert 2s is dependent on the rate at which matter is accreted by the central black hole. They find that those which do not have a hidden broad line region have a mass accretion rate less than $10^{-3}$ Eddington units. Thus, low accretion rate systems may not form a stable broad line region so they would not show broad lines in their spectra, yet will still be unobscured. On the other hand, our multiwavelength analysis has shown a significant number (4 out of 6) Seyfert 2 galaxies in which the absorbing column appears to have been completely underestimated. The distribution of absorbing column densities is a key ingredient in models of the cosmic X-ray background, especially the fraction of Compton thick AGN, which is required to fit the 30 keV peak in the spectral energy distribution \\citep{gilli07}. The derived $N_{\\rm{H}}$ distribution from these models generally over-predicts the number of Compton thick AGN observed by many authors (e.g. \\cite{risaliti99}) and also under-predicts the number of unobscured AGN. Our work potentially provides an answer to this discrepancy, at least for the 12 ${\\umu}$m sample. From our small but well selected sample, we find that 2/3 of Seyfert 2 galaxies with an $N_{\\rm{H}}$ reported to be $<10^{21}$ cm$^{-2}$ should in fact be designated with an $N_{\\rm{H}}>10^{23}$ cm$^{-2}$, and in reality are probably Compton thick. The 12 ${\\umu}$m sample contains 79 Seyfert 2s with good quality X-ray coverage, and of these 19 have been reported to be unabsorbed (24\\%), compared to 4\\% found by \\cite{risaliti99} for their sample of local Seyfert 2s. Combined with the consideration that many unabsorbed Seyfert 2s of the 12 ${\\umu}$m sample have a dubious classification as Seyfert 2, our findings alter the $N_{\\rm{H}}$ distribution of the 12 ${\\umu}$m sample to better agree with those derived from the X-ray background and other observations. Misidentifying these heavily obscured objects due to contamination by the host galaxy, as may be the case in NGC 3486, 3976 and 4501, should be an effect only seen in local AGN, as in deep surveys, only the brightest AGN are selected, where the nucleus will dominate over any galactic emission, but if scattered nuclear emission dominates the soft X-ray then this could be a significant issue even at high redshift." }, "0808/0808.3013.txt": { "abstract": "Extremely red objects, identified in the early {\\it Spitzer} Space Telescope observations of the bright-rimmed globule IC 1396A and photometrically classified as Class I protostars Class II T Tauri stars based on their mid-infrared colors, were observed spectroscopically at 5.5--38 $\\mu$m ({\\it Spitzer} InfraRed Spectrograph), at the 22 GHz water maser frequency (NRAO Green Bank Telescope), and in the optical (Palomar Hale 5-m), to confirm their nature and further elucidate their properties. The sources photometrically identified as Class I, including IC1396A:$\\alpha$, $\\gamma$, $\\delta$, $\\epsilon$, and $\\zeta$, are confirmed as objects dominated by accretion luminosity from dense envelopes, with accretion rates 1--10$\\times 10^{-6} M_{\\odot}$~yr$^{-1}$ and present stellar masses 0.1--2 $M_{\\odot}$. The Class I sources have extremely red continua, still rising at 38 $\\mu$m, with a deep silicate absorption at 9--11 $\\mu$m, weaker silicate absorption around 18 $\\mu$m, and weak ice features including CO$_2$ at 15.2 $\\mu$m and H$_2$O at 6 $\\mu$m. % The ice/silicate absorption ratio in the envelope is exceptionally low for the IC 1396A protostars, compared to those in nearby star-forming regions, suggesting the envelope chemistry is altered by the radiation field or globule pressure. % Only one 22 GHz water maser %(characteristic of a Class 0 protostar) was detected in IC 1396A; it is coincident with a faint mid-infrared source, offset from near the luminous Class I protostar IC 1396A$\\gamma$. The maser source, IC 1396A$\\gamma_{b}$, has luminosity $<0.1 L_{\\odot}$, the first H$_{2}$O maser from such a low-luminosity object. Two near-infrared H$_{2}$ knots on opposite sides of IC 1396A:$\\gamma$ reveal a jet, with axis clearly distinct from the H$_{2}$O maser of IC 1396A:$\\gamma_{b}$. The objects photometrically classified as Class II, including IC1396A:$\\beta$, $\\theta$, 2MASSJ 21364964+5722270, 2MASSJ 21362507+5727502, LkH$\\alpha$ 349c, Tr 37 11-2146 and Tr 37 11-2037, are confirmed as stars with warm, luminous disks, with a silicate emission feature at 9--11 $\\mu$m, and bright H$\\alpha$ emission, so they are young, disk-bearing, classical T Tauri stars. % The disk properties change significantly with source luminosity: low-mass (G--K) stars have prominent 9--11 emission features due to amorphous silicates while higher-mass (A--F) stars have weaker features requiring abundant crystalline silicates. A mineralogical model that fits the wide and low-amplitude silicate feature of IC1396A:$\\theta$ requires small grains of crystalline olivine (11.3 $\\mu$m peak) and another material to to explain the its 9.1 $\\mu$m peak; reasonable fits are obtained with a phyllosilicate, quartz, or relatively large ($>10$ $\\mu$m) amorphous olivine grains. The distribution of Class I sources is concentrated within the molecular globule, while the Class II sources are more widely scattered. Combined with the spectral results, this suggests two phases of star formation, the first (4 Myr ago) leading to the widespread Class II sources and the central O star of IC 1396, and the second ($< 1$ Myr ago) occurring within the globule. The recent phase was likely triggered by the wind and radiation of the central O star of the IC 1396 \\ion{H}{2} region. %with possible additional contributions from the outflows of LkH$\\alpha$ 349a,c %and some nearby B stars. ", "introduction": "{\\it Spitzer} has opened widely the mid-infrared (3.6--24 $\\mu$m) window for studies of star formation. In the earliest observations with the observatory, as in many subsequent ones, previously unknown, mid-infrared-bright sources have been discovered in star-forming regions. In this paper we follow up \\citep[][Paper I]{reachIC} on the mid-infrared sources in the bright-rimmed globule IC 1396A, known as the Elephant Trunk Nebula, a dense globule in the large, nearby (750 pc) \\ion{H}{2} region IC 1396, which is is excited by the 4 Myr-old O6 star HD 206267 \\citep[see][]{weikard}. Based on comparison to measured colors to those of young stellar objects in Taurus \\citep{kenyonhartmann}, we suggested 10 sources may be the very early Class I stage, which is thought to only last $10^{5}$ yr \\citep{spbook}, indicating that the globule is a site of very recent star formation. Detecting new protostars in a wide variety of star forming regions is an important expansion of previous work that by necessity has concentrated on the very closest star forming regions such as Taurus and Orion. While these nearby regions provide excellent opportunities to observe the widest range of star formation (from brown dwarfs to massive stars), they are necessarily parochial. Contrary to a long-standing common belief that the Sun formed in a quiescent region like Taurus, meteoritic evidence indicates the Sun likely formed near a high-mass star, whose \\ion{H}{2} region, winds, and supernova explosion influence the evolution and composition of the protoplanetary disk \\citep{meteorite,tachibana}. To search for present-day star-forming regions similar to those in which the Sun formed, we must consider massive star forming regions. In such regions, stars can be formed by radiative driven implosion of moderately dense cores \\citep{lefloch}, and planet-building disks are exposed to strong ultraviolet radiation \\citep{proplyd}. A major goal of future star formation studies should be to determine the effects of massive stars both in triggering star formation and shaping disks around young stellar objects. In this paper, we present follow-up observations to examine the nature of the sources in IC 1396A in more detail. New {\\it Spitzer} mid-infrared spectra demonstrate that the photometrically extremely red objects are indeed Class I protostars, with dense, highly-optically-thick envelopes shrouding a moderately luminous core inside a dust photosphere. We show that at least one of the extremely red objects, IC 1396 A:$\\gamma$, powers a molecular outflow with H$_2$O maser emission, characteristic of a very young ($\\sim 10^{4}$ yr) Class 0 protostars \\citep{furuya}. The photometrically-classified Class II objects are compared to classical T Tauri stars, based on H$\\alpha$ emission and mid-infrared disk spectra. ", "conclusions": "We studied a mid-infrared-selected sample of young stellar objects from {\\it Spitzer} imaging (Paper I) using spectroscopic observations from radio to optical wavelengths. The photometric classifications in Paper I are confirmed by the spectroscopy. The Class I sources have deep silicate absorption features and the vast majority of their luminosity arises in the infrared. None of them show H$_{2}$O masers (characteristic of Class 0) or optical (Class II) counterparts. The only H$_{2}$O maser associated with our original sample turns out to be a separate object, IC 1396A:$\\gamma$b, with much lower luminosity $\\sim 0.02 L_{\\odot}$ than the protostar we identified originally, IC 1396A$:\\gamma$. The mid-infrared-selected Class I sources can be fitted with 0.1--2 $M_{\\odot}$ stars with ages $<2\\times 10^{5}$ yr. Compared to stars forming in the more quiescent Taurus molecular cloud, the physical conditions of the globule from which the stars are forming include a higher radiation field due to the O6 star exciting the \\ion{H}{2} region and nearby young B stars. There is also a higher pressure due to compression by the ionization front from the O6 star and winds from the many young and still-forming stars. The wind from the $\\sim 5 M_{\\odot}$ star LkH$\\alpha$ 349a has cleared a hole in the center of the dense globule, and outflows from the Spitzer-discovered Class I objects appear to have swept material into `compartments' whose walls are over-pressured. % The spectroscopy results show that the Class I objects are brighter in the far-infrared than the theoretical models predict, suggesting they have a warmer envelope. % Ice features in the Class I object envelopes differ somewhat from those seen around such objects in other star-forming regions. While the same set of features is present, their amplitude is lower, per unit silicate dust absorption, and the wavelengths are slightly shifted. Both of these effects suggest the Class I envelopes are warmer than in Class I sources in other star-forming regions, perhaps due to the warmer gas temperature in the star-forming material in the globule IC 1396A. % The somewhat different physical conditions of the Class I envelopes in this globule will affect the chemistry at the time when planet formation is just beginning. Because the Sun is thought to have formed near an O star, and most nearby star-forming regions lack an O star, astronomical searches for the physical conditions of the early solar nebula would be best conducted in regions like IC 1396. The stars photometrically classified as Class II are confirmed as such based on the presence of optical spectra with bright H$\\alpha$ emission lines and mid-infrared spectra with silicate features in emission. Most of the disk spectra are similar to other commonly-seen T Tauri star spectra, dominated by amorphous silicates. But the star IC 1396A:$\\theta$ has a particularly structured silicate feature with two `horns' at 9.1 and 11.3 $\\mu$m. The 11.3 $\\mu$m horn can be explained by crystalline olivine. The origin of the 9.1 $\\mu$m horn cannot be uniquely identified, with reasonably good fits provided by quartz (SiO$_{2}$), a phyllosilicate (montmorillonite), or $>12$ $\\mu$m radius amorphous olivine grains. Similar spectra are evident in other wide samples of T Tauri stars. Such spectra show that the planet-building material around at least some stars is dominated by crystalline silicates and is significantly different from interstellar silicates. The star formation history in IC 1396A cannot be described as a single ``burst'' of star formation: Class I and II sources in and near the globule span a range of ages and masses in a small region (2 pc across). It appears that the earliest star formation in IC 1396 occurred approximately 5 Myr ago, based on the ages of the Class II sources and the central O6 star. Some stars ahead of the present-day bright rim of IC 1396A have become exposed by the recession of the globule and by their own outflows. Recently, $< 1$ Myr ago, a powerful wind from the young AeBe star LkH$\\alpha$ 349a blew a bubble through the center of the globule. The star formation in IC 1396A we observe now as Class I sources occurred within the last 0.2 Myr, based on the ages of the Class I sources. The most recent star formation occurs preferentially behind the center of the globule, just inside bright rims. The specific trigger for the most recent star formation may have included not only pressurization of the globule by the O star exciting IC 1396, but also the pressure of outflows and jets from the earlier generation of stars and in particular those with powerful winds like LkH$\\alpha$ 349a. The young stars within IC 1396A appear to be reshaping the globule from the inside, clearing `compartments' within the globule. Gradually, the globule is being fragmented due to the exponentially increasing number of stars (if indeed they are sequentially triggered). The mass contained in walls of the compartments will eventually become small enough, and their overlapping structure chaotic enough, that further star formation in the globule will end. The limiting factor in the star formation efficiency of this one globule will be the dynamic effect of energy input from the stars formed within it." }, "0808/0808.3383_arXiv.txt": { "abstract": "% The Ara OB1a association is one of the closest sites where triggered star formation is visible for multiple generations of massive stars. At about 1.3 kpc distance, it contains complex environments including cleared young clusters, embedded infrared clusters, CO clouds with no evidence of star formation, and clouds with evidence of ongoing star formation. In this review we discuss the research on this region spanning the last half-century. It has been proposed that the current configuration is the result of an expanding wave of neutral gas set in motion between 10--40 million years ago in combination with photoionization from the current epoch. ", "introduction": "Ara OB1a, while somewhat enigmatic, may be one of the best examples of triggered star formation in the local Galaxy. The triggering in the most active portion is easily imagined from images of the $<$ 3 Myr NGC~6193/RCW~108-IR complex such as Figure~\\ref{ESO}. However the full association may be as much as 50 Myr in age and cover several degrees on the sky. Ara OB1 was first studied in detail by Whiteoak (1963). He used photographic photometry and objective prism spectroscopy to identify about 35 O and B star members. Whiteoak's work was a follow up to an H$\\alpha$ survey by Rodgers, Campbell \\& Whiteoak (RCW; 1960), which cataloged 181 H$\\alpha$ emission regions in the southern sky, including RCW 108. NGC~6193 is an open cluster, discovered by James Dunlop in 1826, that is dominated by a pair of O stars, HD 150135 and HD 150136. These are the brightest optically revealed O stars in the association and are thought to be responsible for ionizing the bright rim of emission to the west (NGC 6188, discovered by John Herschel in 1836) which separates NGC 6193 from RCW 108--IR. The youth of the region is clear in sky survey plates and Figure~\\ref{ESO} which show concurrent regions of ionized gas and dust lanes. Whiteoak noted two additional clusters in the region, NGC~6204 -- about 2 degrees to the northeast -- and NGC~6167 -- about 1 degree to the southwest. Ara OB1a is a compact association covering about 1 sq. degree around a central cluster -- NGC 6193. It is generally thought to be about 1.3 kpc away and can be equated with RCW~108. There is some confusion in the literature as to what is actually meant by RCW~108. The original definition of Rodgers, Campbell \\& Whiteoak (1960) refers to all the region where H$\\alpha$ nebulosity is detected, for which they give a size of 210\\arcmin x 120\\arcmin centered at ({\\em l,b} = 336.49, --1.48). Straw et al. (1987) used RCW~108--IR to refer to the embedded IR cluster about 15\\arcmin\\ to the west of the O stars in NGC~6193 and which is identified with IRAS~16362--4845. The confusion arises when RCW~108-IR is abbreviated by dropping the \"--IR\". For the remainder of this paper we will refer to the embedded cluster as RCW~108--IR or IRAS~16362--4845. Moffat \\& Vogt (1973) measured photometry for 13 stars within 4\\arcmin\\ of the center of NGC 6193 and found E$_{B-V}$= 0.4 and a distance of 1360 pc. Herbst \\& Havlen (1977) describe Ara OB1 as having a ``diamond ring'' appearance, with RCW 108--IR as the ``diamond'' and a thin circular dust ring making up almost half of the ``ring''. They performed photoelectric and photographic photometry on 702 stars in the region. Herbst (1974, 1975) identified parts of this region as an ``R association'' as there were three early type stars associated with reflection nebulosity. These stars may mark a separate site of star formation within Ara OB1a. We will discuss this further in Section 6. Mid--infrared MSX \\& IRAS maps of the region show several condensations. The brightest is coincident with RCW~108--IR and two apparently related peaks IRAS 16379--4856 and IRAS~16348--4849. % One of the earliest radio studies of the region was in the survey by Shaver \\& Goss (1970). They made a 5 GHz and 408 MHz survey of over 250 Galactic radio source including RCW 108. RCW 108 was remarkable for having one of the smallest emission regions -- unresolved at 3\\arcmin\\ and having a relatively high density and temperature of the electron population. In addition to Ara OB1a, Whiteoak (1963) identified a background O star cluster coincident with NGC~6193 but with a different distance modulus. While the foreground Ara OB1a has a distance modulus of 10.5$\\pm 0.6$ the second group of about 13 O and B stars (Ara OB1b) has a distance modulus of 12.7 $\\pm 0.5$ or about 3500 pc. The central O stars of the two associations are offset by about 2$^{\\rm o}$ along the Galactic plane (Humphreys 1978), so this is a concern for membership determination. \\begin{figure}[!ht] \\begin{center} \\includegraphics[scale = 0.5, angle = 0]{figure1.eps} \\end{center} \\caption{Salient features of the most active portion of Ara OB1a. This image is a B,V H$\\alpha$ (blue, green and red respectively) from the wide field imager on the MPI-ESO 2.2m telescope. The field of view is 32\\arcmin x 32\\arcmin\\ or about 12 pc on a side.} \\label{ESO} \\end{figure} ", "conclusions": "" }, "0808/0808.4089_arXiv.txt": { "abstract": "We have made a new compilation of observations of maximum stellar mass versus cluster membership number from the literature, which we analyse for consistency with the predictions of a simple random drawing hypothesis for stellar mass selection in clusters. Previously, Weidner and Kroupa have suggested that the maximum stellar mass is lower, in low mass clusters, than would be expected on the basis of random drawing, and have pointed out that this could have important implications for steepening the integrated initial mass function of the Galaxy (the IGIMF) at high masses. Our compilation demonstrates how the observed distribution in the plane of maximum stellar mass versus membership number is affected by the method of target selection; in particular, rather low $n$ clusters with large maximum stellar masses are abundant in observational datasets that specifically seek clusters in the environs of high mass stars. Although we do not consider our compilation to be either complete or unbiased, we discuss the method by which such data should be statistically analysed. Our very provisional conclusion is that the data is {\\it not} indicating any striking deviation from the expectations of random drawing. ", "introduction": "It is well known (following \\citet{weidner+kroupa2004,weidner+kroupa2006} and \\citet{oey+clarke2005}) that in the case of clusters containing fewer than $\\sim 100$ OB stars (i.e. those with mass $<$ a few $\\times~10^4~\\Msun$) the maximum stellar mass increases with cluster mass. At higher cluster mass scales, the value of the maximum stellar mass saturates at around 150--200~\\Msun\\ for reasons that are not entirely clear \\citep[see e.g.][]{zinnecker+yorke2007}. In this paper, we restrict ourselves to considering the lower mass regime. In Section 2 we review why the statistics of maximum stellar masses in clusters of various scales can place constraints on high mass star formation in a cluster context and how rather subtle differences in assumed algorithms for cluster building are imprinted on the integrated galactic IMF (the IGIMF). We emphasise that analysis of the statistics of maximum stellar mass versus cluster mass offers the best prospects for an observational determination of whether the IGIMF should be different from that measured in individual clusters \\citep[see e.g.][]{weidner+kroupa2006,elmegreen2006}. We also stress that competing algorithms can only be distinguished through proper statistical analysis of the observed distributions and that selecting algorithms according to how they reproduce the {\\it mean} trend can be misleading. In Section 3 we present a new (but in all likelihood still incomplete and biased) compilation of observational information on this issue and highlight the sensitivity of the distribution obtained to the method of target selection. In Section 4 we discuss the statistical inferences that can be drawn from the current dataset and conclude (Section 5) with an appeal for further observational information to be used in future analyses. ", "conclusions": "We have high-lighted the difficulty in analysing the data contained in Figure \\ref{mmaxntot} owing to difficulties to assigning regions of the diagram where the data is believed to be complete. Nevertheless, our preliminary conclusion is that we are {\\it not} seeing strong evidence for a systematic suppression in maximum stellar mass in small $n$ clusters in addition to that expected on the basis of the statistics of random drawing (see also the complementary analysis of the statistics of isolated stars by \\citet{parker+goodwin2007}, which reached similar conclusions). Indeed, if anything, the feature of Figure \\ref{mmaxntot} that seems to be most discrepant with the random drawing model is the data hole in the range $m_\\mathrm{max} \\sim 3-10 \\Msun, n \\sim 50-100$. We are however aware that this might indeed be filled in if we have under-estimated the incompleteness in smaller $n$ clusters (particularly due to the effects of dynamical evolution). Our conclusion (in support of the random drawing hypothesis) remains provisional. Although we have set out what we believe to be a statistically correct methodology for analysing the problem, we are highly aware of the difficulties of properly quantifying observational selection effects. We therefore seek further input from observers in compiling a good sample for this kind of analysis." }, "0808/0808.4040_arXiv.txt": { "abstract": "HIZOA J0836-43 is an extreme gas-rich ($M_{\\rm{HI}}$=7.5$\\times10^{10}\\, M_{\\sun}$) disk galaxy which lies hidden behind the strongly obscuring Vela region of the Milky Way. Utilizing observations from the {\\it Spitzer Space Telescope}, we have found it to be a luminous infrared starburst galaxy with a star formation rate of $\\sim 21\\, M_{\\sun}\\, \\rm{yr^{-1}}$, arising from exceptionally strong molecular PAH emission ($L_{7.7\\micron} = 1.50 \\times 10^{9} L_{\\odot}$) and far-infrared emission from cold dust. The galaxy exhibits a weak mid-infrared continuum compared to other starforming galaxies and U/LIRGs. This relative lack of emission from small grains suggests atypical interstellar medium conditions compared to other starbursts. We do not detect significant $[$Ne\\,{\\sc v}$]$ or $[$O\\,{\\sc iv}$]$, which implies an absent or very weak AGN. The galaxy possesses a prominent bulge of evolved stars and a stellar mass of 4.4($\\pm$1.4)$\\times10^{10}\\, M_{\\sun}$. With its plentiful gas supply and current star formation rate, a doubling of stellar mass would occur on a timescale of $\\sim$2 Gyr. Compared to local galaxies, HIZOA J0836-43 appears to be a ``scaled-up\" spiral undergoing inside-out formation, possibly resembling stellar disk building processes at intermediate redshifts. ", "introduction": "Understanding the fundamental origin and formation of galaxies requires the synthesis of multi-wavelength observations and cosmological-based numerical simulations. Advances in both areas have created a compelling view of a cold dark matter dominated universe where structure forms hierarchically \\citep{Wh78}. Of fundamental importance to this formalism is insight into the formation and evolution of galaxy disks \\citep{Str07}. Massive, gas-rich disk galaxies ($M_{\\rm{HI}} > 10^{10} M_{\\sun}$) are considered indicative of relatively unadvanced star building systems, making them ideal laboratories for testing this formalism. However, such systems are rare in the local universe and nearly all massive \\HI-rich disk galaxies are inactive or only passively forming stars \\citep{Spray95}, thus providing few clues as to their formation and evolution. For example, Malin 1, an extreme case of a giant low surface brightness galaxy \\citep{Imp89}, has a dormant star formation rate of $\\sim 0.38\\, M_{\\sun}\\, \\rm{yr^{-1}}$ \\citep{Rah07}. In this letter, however, we present evidence of an \\HI-massive disk galaxy, HIZOA J0836-43, undergoing a vigorous starburst. This galaxy, discovered as part of a blind \\HI\\ survey of the southern Zone of Avoidance, contains 7.5$\\times10^{10}\\, M_{\\sun}$ of \\HI\\ gas, has a total dynamical mass of 1.4$\\times10^{12}\\, M_{\\odot}$ and a 20-cm derived star formation rate of $\\sim$35\\,$M_{\\sun}\\, \\rm{yr^{-1}}$ \\citep{Don06}. It also appears to have a prominent bulge in the near-infrared -- bulge-to-disk ratio of $\\sim 0.80$ in the $K_{s}$ band -- central to an enormous, rapidly rotating \\HI\\ disk. The galaxy is located at $l$=262.48$\\degr$, $b$=-1.64$\\degr$, lying behind the Vela Supernova Remnant of the Milky Way. At optical wavelengths it is largely hidden by foreground gas and dust. We have therefore conducted a detailed infrared study of HIZOA J0836-43, using imaging and spectroscopy from the {\\it Spitzer Space Telescope}, to fully reveal its morphology and past and present evolutionary state. Here we highlight key results from this study which provide new insights for understanding massive galaxy formation and evolution. The adopted distance of HIZOA J0836-43, $D_{L}$=148 Mpc, is from \\citet{Don06}, as derived from its recessional velocity, $v_{hel}$=10689 km\\,$\\rm{s^{-1}}$. ", "conclusions": "For a galaxy in the local universe HIZOA J0836-43 possesses a number of unusual infrared properties; considered in combination with the massive reservoir of gas that feeds it, this could be a rare instance of a local galaxy undergoing inside-out evolution, possibly resembling galaxy formation at earlier epochs. Here we compare properties of HIZOA J0836-43 with those of local and intermediate redshift samples. The paucity of warm dust is evident from the weakly rising continuum seen in its spectrum (Fig. 2b), consistent with its low $L_{24\\micron}/L_{70\\micron}$ colour compared to normal and star-forming systems, e.g., as compared to both the SINGS ({\\it Spitzer} Infrared Nearby Galaxy Survey) sample, and the relatively nearby Great Observatories All-sky LIRG Survey (GOALS). In contrast, the strength of the PAH emission in HIZOA J0836-43, both in luminosity and relative strength of the bands compared to the continuum, is amongst the largest observed in any star-forming galaxy \\citep{Peet04}, implying unusually strong PDR emission. These unusual MIR properties may arise from (1) a paucity of very small grains (VSGs) and/or (2) a soft UV radiation field. Powered by massive star formation, transiently heated VSGs are thought to be the source of MIR radiation. A weaker radiation field would give rise to cooler VSGs \\citep{Dra07}, which would re-radiate at longer FIR wavelengths along with the large grain, cool (T $\\sim$20 K) component. Both mechanisms would be consistent with an evolutionary phase in which the activity is confined to heavily gas/dust-obscured star forming regions, thus shielding the hard UV radiation from the rest of the disk, which in turn would yet to have experienced significant grain processing. \\begin{figure}[!t] \\begin{center} \\includegraphics[width=8.5cm]{f3.eps} \\caption[Comparison with SINGS and Malin 1] {\\small{Infrared luminosity vs. \\HI\\ mass, comparing HIZOA J0836-43, Malin 1 and the SINGS galaxy sample. For Malin 1 (yellow triangle) a 70\\micron\\ upper limit \\citep{Rah07} is shown; upper limits for \\HI\\ mass are shown as black triangles.} } \\label{fig:3} \\end{center} \\end{figure} The growth progress of HIZOA J0836-43 is deduced from the stellar bulge population and current star formation rate. We estimate a stellar mass of 4.4($\\pm$1.4)$\\times10^{10}\\, M_{\\sun}$ for the galaxy \\citep[from the relation of][]{Bell03} and hence a specific star formation rate, SFR per stellar mass, of $\\sim$0.5 $\\rm{Gyr}^{-1}$. Compared to local LIRGS \\citep[see Fig. 5 of][]{Wan06}, this implies active stellar building as facilitated by its plentiful supply of gas, with a doubling of stellar mass in $\\sim$2 Gyr. The specific star formation rate of the galaxy in combination with its stellar mass appears typical for star-forming galaxies at $z\\sim 0.7$, where gas fractions of disks were likely higher compared to local galaxies \\citep{Bell05, Per05}. Fig. 3 shows that the MIR luminosity is strongly correlated with \\HI\\ content; even with its extreme \\HI\\ mass, HIZOA J0836-43 appears consistent with being a ``scaled-up\" disk galaxy, unlike Malin 1 which is explicitly quiescent by comparison. This suggests relatively ``normal'' evolution in HIZOA J0836-43, despite lying at the extreme high end (i.e., early evolutionary stage) of the relation. HIZOA J0836-43 is a gas-rich spiral galaxy exhibiting a warp in its disk (Fig. 1a), likely the result of a disturbance in its recent past ($<<$ 1 Gyr). Such an event could cause the observed starburst as gas from the extended \\HI\\ disk flows into the central region of the galaxy. Hence the starburst is powered by gas consumption, as opposed to a major merger event. Gas-rich galaxies, like HIZOA J0836-43, were likely more common in the distant universe and there is evidence that gas consumption and not merger interactions were driving stellar mass growth in the distant universe \\citep{Dad08}. Recent work has suggested that many LIRGs seen at intermediate redshifts ($z\\sim0.8$) achieve heightened star formation as a result of the high gas fractions of their disks and were less dependent on major interactions, compared to local LIRGs, to induce starburst activity \\citep{Mel08, Mar06}. Observational evidence suggests that disk galaxies evolve along the stellar mass-radius relation and have built stellar mass intensely since $z=1$, on average by means of inside-out growth \\citep{Bar05,Tru06}. Comparing the extended regions of active star formation with the more centrally concentrated stellar bulge distribution (e.g., Fig 1b radial profile) suggests that HIZOA J0836-43 is undergoing vigorous disk building, an instance of inside-out growth. This combined with its PDR-dominated emission manifested as strong PAH emission coupled with a weak MIR continuum, makes it enigmatic in the local universe. Observing a galaxy at such a key point in its evolution could have far reaching implications for theories of galaxy formation and evolution." }, "0808/0808.2046_arXiv.txt": { "abstract": "We discuss the relationship between a symmetry in the neutrino flavour evolution equations and neutrino flavour oscillations in the collective precession mode. This collective precession mode can give rise to spectral swaps (splits) when conditions can be approximated as homogeneous and isotropic. Multi-angle numerical simulations of supernova neutrino flavour transformation show that when this approximation breaks down, non-collective neutrino oscillation modes decohere kinematically, but the collective precession mode still is expected to stand out. We provide a criterion for significant flavour transformation to occur if neutrinos participate in a collective precession mode. This criterion can be used to understand the suppression of collective neutrino oscillations in anisotropic environments in the presence of a high matter density. This criterion is also useful in understanding the breakdown of the collective precession mode when neutrino densities are small. ", "introduction": "Because of neutrino-neutrino forward scattering or neutrino self-interaction \\cite{Fuller:1987aa,Notzold:1988kx,Pantaleone:1992xh} neutrinos can experience collective flavour transformation in environments such as the early Universe (e.g., \\cite{Samuel:1993uw,Kostelecky:1994dt,Samuel:1995ri,Pastor:2001iu,Dolgov:2002ab,Abazajian:2002qx}) and supernovae (e.g., \\cite{Qian:1994wh,Pastor:2002we,Balantekin:2004ug,Fuller:2005ae}) where neutrino number densities can be very large. This phenomenon is different from the conventional Mikheyev-Smirnov-Wolfenstein (MSW) effect \\cite{Wolfenstein:1977ue,Mikheyev:1985aa} in that the flavour evolution histories of neutrinos in collective oscillations are coupled together and must be solved simultaneously. The possibility and consequences of collective neutrino oscillations in supernovae were not well appreciated until the discovery that the ordinary matter can be ``ignored'' in such phenomena \\cite{Duan:2005cp} and the first numerical demonstrations of ``stepwise spectral swapping'' (or ``spectral split'') \\cite{Duan:2006jv,Duan:2006an} which is the imprint left by the collective flavour transformation on neutrino energy spectra. Significant progress has been made towards understanding collective neutrino oscillations in supernovae (see, e.g., \\cite{Duan:2009cd} for a brief review and references therein). In particular, Raffelt and Smirnov \\cite{Raffelt:2007cb} demonstrated an adiabatic (precession) solution for the homogeneous, isotropic neutrino gas. This solution has been shown to agree in part with the results of ``single-angle'' simulations of supernova neutrino oscillations \\cite{Duan:2007fw}. These single-angle simulations essentially neglect the anisotropic nature of the supernova environment by assuming that the flavour evolution histories of neutrinos along all trajectories are identical to those along a ``representative'' trajectory, usually taken to be the radial trajectory \\cite{Qian:1994wh,Dasgupta:2008cu}. The adiabatic precession solution requires that at any time all neutrinos reside in a pure collective oscillation mode, the ``precession'' mode, which, as shown by Duan \\etal \\cite{Duan:2006an,Duan:2007mv}, would explain the spectral swap phenomenon in the single-angle simulations. However, the real supernova environment is highly inhomogeneous and anisotropic. Here by ``inhomogeneous'' and ``anisotropic'' we are referring to the neutrino fields. Of course, the matter density distributions in the supernova environment are also likely to be inhomogeneous and anisotropic. To date there are a few ``multi-angle'' simulations \\cite{Duan:2006jv,Duan:2006an,Duan:2007bt,Fogli:2007bk} which, like the single-angle calculations, also adopt spherically symmetric supernova models but do treat flavour evolution along different neutrino trajectories in a self-consistent way. These multi-angle calculations also exhibit spectral swaps. It is still not understood how the spectral swap phenomenon arises in the (anisotropic) multi-angle context. In fact, some studies seem to suggest that collective neutrino oscillations in the isotropic and anisotropic environments can be very different. For example, collective neutrino oscillations of the bipolar type can experience ``kinematic decoherence'' and be disrupted in anisotropic environments \\cite{Raffelt:2007yz,EstebanPretel:2007ec}. Additionally, a very large matter background can result in neutrino oscillation phase differences between different neutrino trajectories in an anisotropic neutrino gas \\cite{Qian:1994wh}, and this effect recently has been shown to result in suppression of collective neutrino oscillations \\cite{EstebanPretel:2008ni}. In this paper we discuss a SU($N_\\mathrm{f}$) rotation symmetry in the neutrino flavour evolution equations, where $N_\\mathrm{f}=2$ and 3 for the two-flavour and three-flavour neutrino mixing schemes, respectively. The collective precession mode for neutrino oscillations can ensue from this symmetry, even in inhomogeneous, anisotropic environments. This result explains the puzzling observations of the spectral swapping phenomenon in both the single-angle and multi-angle simulations of supernova neutrino oscillations. The rest of this paper is organized as follows. In \\sref{sec:framework} we lay out the general framework for neutrino flavour transformation and discuss the SU($N_\\mathrm{f}$) rotation symmetry in the flavour evolution equations for a dense neutrino gas. In \\sref{sec:ap-sol} we show how the collective precession mode for neutrino oscillations can arise from this symmetry in various environments. We also give criteria for when the collection precession mode can occur. In \\sref{sec:sn} we present a new multi-angle simulation of supernova neutrino oscillations. We analyze the results of this calculation guided by our understanding of the collective precession mode. In \\sref{sec:conclusions} we give our conclusions. ", "conclusions": "} We have demonstrated the existence of the SU($N_\\mathrm{f}$) rotation symmetry in the neutrino flavour evolution equations. This symmetry can facilitate the establishment of the collective precession mode for neutrino flavour oscillations in various environments. The stepwise-spectral-swapping phenomenon can develop from such a collective neutrino oscillation mode in, e.g., supernovae. We have also given criteria for significant neutrino flavour oscillations to occur if neutrinos are entrained in the collective precession mode. These criteria can be used to understand the suppression of collective neutrino oscillations in anisotropic environments in the presence of a large matter density \\cite{EstebanPretel:2008ni}. These criteria also illuminate the process of the breakdown of collective oscillations when neutrino densities are low. The results obtained in both our new simulation and previous multi-angle calculations for supernova neutrino oscillations can be understood in terms of the collective precession mode and the criteria we have provided. There remains much to be learned about the collective precession mode for neutrino oscillations. This collective mode can not exist when there is no solution to \\eref{eq:prec-ansatz-var} and \\eref{eq:L-cons-var}. A related interesting observation is that collective oscillations, if any, for a symmetric system with equal numbers of neutrinos and antineutrinos are quickly disrupted in the presence of even an infinitesimal anisotropy and flavour equipartition is obtained as a result \\cite{Raffelt:2007yz}. In the flavour pendulum analogy \\cite{Hannestad:2006nj}, the asymmetry of neutrinos and antineutrinos constitutes the internal spin of the flavour pendulum. The existence of this internal spin causes the flavour pendulum to undergo the precession which represents the collective precession mode for neutrino oscillations. There exists no collective precession mode in a symmetric neutrino system --- this explains the finding in \\cite{Raffelt:2007yz}. Meanwhile, it has been shown in \\cite{EstebanPretel:2007ec} that a typical asymmetry in the neutrino and antineutrino fluxes in the supernova environment will suppress multi-angle decoherence and, therefore, make the collective precession mode for neutrino oscillations possible. \\ack This work was supported in part by DOE grants DE-FG02-00ER41132 at the INT, DE-FG02-87ER40328 at UMN, NSF grant PHY-06-53626 at UCSD, and an IGPP/LANL mini-grant. This research used resources of the National Energy Research Scientific Computing Center, which is supported by the Office of Science of the US Department of Energy under Contract No.\\ DE-AC02-05CH11231. We thank J Carlson, J J Cherry, A Friedland, W Haxton, D B Kaplan, C Kishimoto, C Lunardini, A Mezzacappa, G Raffelt and P Romatschke for insightful discussions." }, "0808/0808.0619_arXiv.txt": { "abstract": "{}{We present a panchromatic study, involving a multiple technique approach, of the circumstellar disc surrounding the T~Tauri star IM~Lupi (Sz~82).} {We have undertaken a comprehensive observational study of IM Lupi using photometry, spectroscopy, millimetre interferometry and multi-wavelength imaging. For the first time, the disc is resolved from optical and near-infrared wavelengths in scattered light, to the millimetre regime in thermal emission. Our data-set, in conjunction with existing photometric data, provides an extensive coverage of the spectral energy distribution, including a detailed spectrum of the silicate emission bands. We have performed a simultaneous modelling of the various observations, using the radiative transfer code MCFOST, and analysed a grid of models over a large fraction of the parameter space via Bayesian inference. }{We have constructed a model that can reproduce all of the observations of the disc. Our analysis illustrates the importance of combining a wide range of observations in order to fully constrain the disc model, with each observation providing a strong constraint only on some aspects of the disc structure and dust content. Quantitative evidence of dust evolution in the disc is obtained: grain growth up to millimetre-sized particles, vertical stratification of dust grains with micrometric grains close to the disc surface and larger grains which have settled towards the disc midplane, and possibly the formation of fluffy aggregates and/or ice mantles around grains.}{} ", "introduction": "During the first stages of planet formation following the core nucleated accretion scenario \\citep{Lissauer06PPV}, evolution of dust grains within the protoplanetary disc surrounding the central forming object is expected. Micrometre size dust grains will start to grow by coagulation during low relative velocity collisions, leading to the formation of larger, potentially fluffy aggregates \\citep[][and references therein]{Beckwith00,Dominik06PPV,Natta06PPV} that will ultimately give birth to kilometre size planetesimals promoting gravitational focusing and eventually leading to planets. Timescales for the formation of these aggregates strongly depend on the physics of aggregation as well as the physical conditions within the disc, such as differential velocities between grains and grain-gas interactions. \\begin{figure*} \\includegraphics[height=0.32\\hsize]{sz82_f606nup_sqrt.eps}% \\hspace{\\stretch{0.5}}% \\includegraphics[height=0.32\\hsize]{sz82_f814nup_sqrt.eps}% \\hspace{\\stretch{0.5}}% \\includegraphics[height=0.32\\hsize]{IMLup_1.6.eps}% \\caption{ The IM Lupi circumstellar disc observed in scattered light by HST. {\\bf Left and middle panels:} respectively F606W and F814 \\emph{WFPC2} PC1 PSF-subtracted images. The dark diagonal feature running through the star is an artifact of charge bleeding along the CCD detector columns. {\\bf Right panel:} \\emph{NICMOS} PSF-subtracted coronagraphy at 1.6 $\\mu$m. The central circle represents the 0.3\\arcsec\\ radius coronagraphic obscuration. All images are shown in square root stretch, with North up and East to the left. The field of view is 5x5\\arcsec (dashed boxed in Fig.~\\ref{fig:schema}). \\label{fig:obs_HST}} \\end{figure*} \\begin{figure} \\centering \\includegraphics[width=\\hsize]{IM_Lup_schema3.eps} \\caption{\\emph{NICMOS} F160W image in $\\log$ stretch. The field of view is 11$\\times$11\\arcsec\\ with North up and East to the left. The dashed box corresponds to the field of view in Fig.~\\ref{fig:obs_HST}. The central white circle represents the 0.3\\arcsec\\ radius coronagraphic obscuration. The full green lines indicate the upper and lower scattering surfaces of the disc (separated by the dark lane which corresponds to the disc midplane). The green dashed circle represents the possible envelope surrounding the disc.\\label{fig:schema}} \\end{figure} In parallel to grain growth, dust grains are expected to settle towards the disc midplane and then to migrate inward as a result of the conjugate actions of the stellar gravity and gas drag \\citep[e.g.][]{Barriere05,Fromang06}. This settling is highly dependent on the grain size. Small grains ($<$~1~$\\mu$m) are strongly coupled to the gas, they follow its motion, and do not settle at all. Conversely, grains in the mm-cm regime are considerably slowed down by the gas drag. They completely decouple and settle very rapidly in a thin midplane. The result is the formation of a dust sub-disc close to the equatorial plane \\citep[e.g.,][]{Safronov69,Dubrulle95} that may become unstable and produce planetesimals when the density of the dust layer exceeds the gas one \\citep[e.g.,][]{Goldreich73,Schrapler04,Johansen07}. Models suggest grain growth to large particle sizes with a population of small grains remaining close to the surface and larger grains deeper in the disc \\citep[e.g.][]{Dominik06PPV}. However, many details of process remain uncertain. For example, it is not clear how dust grains overcome the ``metre size barrier'' without being accreted onto the star \\citep[e.g.\\ ][]{Weidenschilling77,Brauer08} or destroyed by high speed collisions \\citep[e.g.][]{Jones96,Blum00}, or how Kelvin-Helmholtz instabilities prevent the dust sublayer from fragmenting \\citep[e.g.][]{Johansen06}. \\defcitealias{Dullemond04}{D04} \\def\\D04{\\citetalias{Dullemond04}} Detection of the large building blocks of planets (particle sizes $> 1$m) in the disc midplane is far beyond current observational capacities, but we can expect to detect signatures of their formation. The stratified structure resulting from grain growth and settling has direct consequences on disc observables like their spectral energy distributions (SEDs) and scattered light images \\citep[e.g.][hereafter D04]{Dullemond04}. A variety of observational techniques have been used to obtain insights into the disc properties and their dust content: SEDs \\citep[e.g.][]{D'Alessio01}, scattered light images \\citep{Watson06PPV}, thermal emission maps in the millimetre regime \\citep{Dutrey06PPV}, mid-infrared spectroscopy \\citep[e.g.][]{Kessler-Silacci06}. However, each technique only provides a limited view of a disc. SED analysis leads to multiple ambiguities (e.g., \\citealp{Thamm94,Chiang01}) since the lack of spatial resolution precludes from solving the degeneracies between model parameters like geometry and opacity. Spatially resolved observations in a single spectral band (scattered light in the optical or near-infrared regime or thermal emission in the millimetre domain) also gives incomplete information about the disc. As the dust opacity is a steep function of wavelength and because high temperature gradients are present within the disc, a single-band observation gives insight to only a limited region of the disc. Thus, scattered light images probe the surface layers of these optically thick discs at large radii whereas, millimetre observations are mainly sensitive to the bulk of the disc mass closer to the midplane. To precisely study fine physical processes like dust evolution and stratification, it is necessary to combine the aforementioned methods in a multi-wavelength, multiple technique observational and modelling approach. In this paper, we study the circumstellar environment of the T~Tauri star IM~Lupi. The disc surrounding IM~Lupi (Schwartz~82, HBC~605, IRAS~15528-3747) was first detected in scattered light with the {\\it WFPC2} instrument on board the \\emph{Hubble Space Telescope} (\\emph{HST}) as part of a T~Tauri star imaging survey (Stapelfeldt et al. 2008, in prep.). IM~Lupi is an M0 T~Tauri star located within the Lupus star forming clouds. It is one of four young stellar objects in the small $^{13}$CO(1-0) Lupus~2 core near the extreme T~Tauri star RU~Lupi \\citep{Tachihara96}. Despite the low accretion-related spectroscopic activity of IM~Lupi \\citep{Reipurth96,Whichmann99}, longer wavelength observations reveal ample evidence for circumstellar material in the system with strong mm continuum emission \\citep{Nuernberger97}. Single dish $^{12}$CO and $^{13}$CO line observations indicate that the disc is gas rich and are consistent with a rotating disc model \\citep{vanKempen07}. The 3.3\\,mm continuum emission from the disc was spatially resolved by \\cite{Lommen07}. Preliminary models by \\cite{Padgett06} showed that the infrared excess is well-reproduced by a disc model. We have built a rich data set of observations of the disc: a well sampled SED, \\emph{HST} multiple wavelength scattered light images, \\emph{Spitzer} near- and mid-infrared spectroscopy and \\emph{SMA} millimetre emission maps. These observations provide different, complementary detailed views of the disc structure and dust properties. We investigate whether all of these observations can be interpreted in the framework of a single model from optical to millimetre wavelengths, and whether we can derive quantitative conclusions regarding the evolutionary stage of the disc. In section~\\ref{sec:obs}, we describe the observations and data reduction procedure. In section~\\ref{sec:simple_estimation}, we first draw constraints on the disc and dust properties from the various observations and then, in section~\\ref{sec:grid_modelling}, we present simultaneous modelling of these observations. A detailed study of the dust properties is presented in section~\\ref{sec:dust_properties}. In section~\\ref{sec:discussion}, we discuss the implication of the results, and section~\\ref{sec:conclusion} contains the concluding remarks. ", "conclusions": "\\label{sec:discussion} \\subsection{A border line CTTS but with a massive disc} IM~Lupi displays only a modest amount of emission-line activity, with a H$\\alpha$ equivalent width which is known to vary from 7.5 to 21.5\\,\\AA\\ \\citep{Batalha93}. Studies of H$\\alpha$ emission show evidence of accretion, including variability: \\cite{Reipurth96} concluded that the H$\\alpha$ feature shows an inverse P Cygni profile (classification IV-Rm) and \\cite{Whichmann99} observed a III-R profile. \\emph{International Ultraviolet Explorer} low dispersion LWP spectra show that the Mg\\,II 2\\,798\\,\\AA\\ line also varies, by a factor of about 2 in net flux \\citep{Valenti03}. The relatively weak emission lines and lack of optical veiling caused \\cite{Finkenzeller87} and \\cite{Martin94} to classify this object as a weak-line T~Tauri star, although our results show it would be more properly categorised as a borderline classical T Tauri star. IM~Lupi is a relatively weak-lined T Tauri star with a large and massive circumstellar disc. Our modelling indicates a large dust disc mass of $M_\\mathrm{dust} = 10^{-3}\\,M_\\odot$, extending up to a radius of 400\\,AU. This large mass is puzzling given the weakness of the H$\\alpha$ line, which suggests a low accretion rate. It is however possible that the diagnostics of accretion have only been observed during periods of low or moderate accretion. \\subsection{Disc structure \\label{sec:disc_structure}} Our multi-technique modelling allows us to quantitatively constrain most of the geometrical parameters of the disc. A flared geometry with a scale height of 10\\,AU at a reference radius of 100\\,AU is required. The best model has a midplane temperature of 14\\,K at 100\\,AU. If we assume the disc is vertically isothermal (with the temperature equal to the midplane temperature), the hydrostatic scale height $\\sqrt{k_B r^3 T(r) / G M_* \\mu}$ (where $\\mu$ is the mean molecular weight) is 12 and 8.5\\,AU, respectively, for a central mass object of 0.5 and 1\\,M$_\\odot$, respectively. This is in good agreement with the scale height deduced from the observations (10\\,AU), indicating that the outer parts of the disc are very likely to be in hydrostatic equilibrium. Figure~\\ref{fig:temperature} presents the calculated midplane temperature compared to the temperature corresponding to the gas scale height of the best model. The agreement is very good at the inner edge and in the outer parts of the disc. In the central parts of the disc ($<30$\\,AU and excluding the very inner edge), the midplane temperature obtained in our passive disc model is too low to explain the scale height required by observations, indicating that an additional heating mechanism (like viscous heating) may be at play in the disc. \\begin{figure} \\includegraphics[width=\\hsize]{temperature.eps} \\caption{Temperature structure of the disc. The full and dashed lines represent the midplane and surface temperatures, respectively. The disc surface at a given radius is defined as the altitude above the midplane where the temperature is maximal. Both temperatures become equal at the inner edge of the disc because it is directly heated by the stellar radiation. The shaded area represents the temperature corresponding to the vertical gas scale height of the best model, assuming that the disc is vertically isothermal and that the central mass is between 0.5 and 1\\,M$_\\odot$. \\label{fig:temperature}} \\end{figure} The inner radius is constrained to be between 0.25 and 0.40\\,AU, which corresponds to a maximum dust temperature around 1\\,000\\,K. The modelling of the SED alone allows values of the inner radius down to 0.15\\,AU corresponding to temperatures close to the dust sublimation temperature (1\\,500\\,K). The modelling of other observations gives additional constraints on the disc parameters and because of the correlations between parameters, this reduces the range of possible values for the inner radius. However, scattered images and millimetre visibilities constrain the large scale structure of the disc. Because we assume that the disc can be described in terms of power-laws from the inner edge to the outer radius, these constraints affect the derived inner radius. However, it is very likely that the description of the disc using power-laws is too simplistic, and more complex geometries may slightly shift the inner radius inwards, making it compatible with the dust sublimation radius. The surface density exponent is found to be close to $-1$. This value corresponds to the median measurement for discs in Taurus-Aurigae obtained by \\cite{Andrews07}. This also corresponds to the theoretical value for a disc in steady-state accretion \\citep{Hartmann98}. The surface density at 5\\,AU is 70\\,g.cm$^{-2}$. This value is within the broad peak around the median value of 14\\,g.cm$^{-2}$ for Taurus \\citep{Andrews07}. Considering a probable stellar mass of $\\approx 1\\,M_\\odot$ and a gas to dust mass ratio of 100, the disc to star mass ratio is $\\approx$ 0.1, meaning the disc may be unstable through gravitational collapse. The stability of the disc will depend on the surface density. Our derived value of the surface density exponent $\\alpha$=-1, as opposed to values of 0 or -2, provides disc stability at all radii according to the Toomre stability criterion \\citep{Toomre64}. Collapse of gravitationally unstable discs \\citep{Durisen07PPV} is one suggested mode for planet formation. The disc of IM Lupi representing about 1/10 of the star mass, local enhancement of density may be sufficient to start planet formation in the disc following this process. \\cite{Hughes08} have shown that simultaneous studies of the dust continuum and CO emissions in several well-studied discs can be reproduced with disc models that include a tapered exponential outer edge and not a sharp outer radius as we have used here. Current observations of IM Lupi do not allow us to study in details the outer edge of the disc but the \\emph{NICMOS} image indicates that dust grains are still present at radii larger than 400\\,AU. However, as the counter nebulae of the disc is seen in scattered light images at 0.8 and 1.6\\,$\\mu$m, we can get a rough estimate of the maximum value for the surface density in front of this second nebulae. Dust present at radii larger than 400\\,AU must be optically thin, allowing us to see the counter nebulae through the disc and the tentative envelope. This can be translated into a upper limit for the surface density of the disc\\footnote{The constraint is stronger at 0.8\\,$\\mu\\mathrm{m}$ than at 1.6\\,$\\mu\\mathrm{m}$, due to the larger dust opacity.}: $\\Sigma (r\\approx500\\,AU) \\times \\kappa_{0.8\\mu\\mathrm{m}} \\lesssim 1$. If we assume that the dust composition and grain size distribution are the same as in the rest of the disc, this corresponds to $\\Sigma (r\\approx500\\,AU) \\lesssim 0.2\\,$g.cm$^{-2}$ at 500\\,AU, \\emph{i.e.} about 3.5 times smaller than the extrapolated density from our disc model, assuming it extends at radii larger than 400\\,AU. This implies that the density at radii larger than 400\\,AU must decrease significantly faster than the $1/r$ dependence we found for the disc in regions inside 400\\,AU. \\subsection{Grain growth and dust settling\\label{sec:dust_evolution}} As with many other classical T Tauri stars, the slope of IM Lupi's millimetre continuum suggests a dust opacity following a law close to $\\kappa_\\mathrm{abs}(\\lambda) \\propto \\lambda^{-1}$, indicating dust grains in the disc are larger than those in the interstellar medium. Millimetre photometry mostly traces the thermal emission of cold dust in the outer part ($> 15$\\,AU) of the disc midplane (see figure~\\ref{fig:flux_cumule}), while the hot grains in the central regions contribute little to this emission. The strong silicate features in the mid-IR spectrum indicate the presence of micron-sized grains in the disc surface inside the first central AU. The silicate emission indeed comes from the central parts of the disc: 90\\,\\% of the emission at 10 and 20\\,$\\mu$m originates from radii smaller than 1 and 10\\,AU, respectively (Figure~\\ref{fig:flux_cumule}). Larger grains ($\\gtrsim 3\\,\\mu$m) appear to be almost absent in these regions (section \\ref{sec:mineralogie}). \\begin{figure} \\centering \\includegraphics[width=0.944\\hsize]{flux_cumule.eps} \\caption{Cumulative received fluxes for the best model as a function of the distance from the star at a wavelength of 10\\,$\\mu$m (full line), 20\\,$\\mu$m (dashed line), 70\\,$\\mu$m (dotted line) and 1.3\\,mm (dot-dash line).\\label{fig:flux_cumule}} \\end{figure} As already mentioned, the slope of the mm continuum and the mid-IR silicate features indicate a stratified structure for the disc, with large grains in the deeper parts of the disc and small grains ($\\lesssim 1\\,\\mu$m) in the surface of the disc. This surface layer of small grains remains difficult to characterise precisely. We have explored a structure where stratification of grain size is vertical and caused by settling of larger particles towards the disc midplane. Our model with vertical settling allows us to accurately reproduce the SED, scattered light images and millimetre emission maps. However, we cannot assess the uniqueness of the solution and other explanations may be envisaged. The warm region emitting the 10 micron silicate feature is close to the star, while the mm/submm continuum comes from the whole volume of the disc. We may be observing small particles close in and big grains in the outer regions of the disc. However, most physical processes: grain growth, radial drift by gas drag in the disc midplane, radiation pressure or stellar wind, will preferentially result in larger grains in the central regions and/or remove small grains from these regions. Mid-infrared interferometric observations (\\citealp{vanBoekel04} for HAeBe stars and \\citealp{Ratzka07} and \\citealp{Schegerer08} for the T~Tauri stars TW~Hydra and RY~Tau respectively) have confirmed these trends by showing the presence of larger and more processed grains at small spatial scales. Timescales for the production of large grains are significantly shorter in the central parts of the disc and it is difficult to imagine a physical process that will efficiently produce large grains in the outer disc without also producing them in the inner disc. Even if the current observations do not allow us to firmly conclude this point, vertical grain size stratification, probably resulting from a combination of settling and enhanced grain growth close to the midplane, appears to be a more natural explanation, and is, for now, our preferred model for the disc of IM~Lupi. Following the previously described argument, the presence of millimetre grains at large radii strongly suggests that such large grains are also present in the central AU, where the process of grain growth should be more efficient. The 10\\,$\\mu$m features, dominated by the emission from grains of size around 1.5\\,$\\mu$m, indicate that the mixing in the disc is sufficient to maintain micrometric grains in the disc surface. Moreover, the absence of 3\\,$\\mu$m grains in the \\emph{IRS} spectrum indicates that the decoupling between gas and dust starts for a grain size between 1 and a few $\\mu$m, \\emph{i.e.} grains larger than this threshold are settled below the surface $\\tau_{10\\,\\mu\\mathrm{m}}$ = 1. Given the high densities in the inner regions of the disc, an increase of 1 in optical depth is obtained over a very small spatial scale. Even a moderate amount of dust settling is sufficient to produce the observed effect. The very low values we derive for the settling index confirm that the settling remains efficiently counterbalanced by vertical mixing, due to turbulence for instance. It is very likely that the process of dust settling evolves with different timescales as a function of the distance from the star. The silicate emission bands provide strong indications of the efficiency of dust settling as a mean of removing grains larger than a few microns from the upper layers in the central parts of the disc. The SED also gives some insights on the presence of settling in the outer parts. The \\emph{MIPS} far-IR fluxes, which are low compared to the mid-IR emission, indicate that the disc intercepts a relatively small fraction of the stellar radiation at large radii ($> 10$\\,AU) compared to the fraction intercepted at radii $< 10$\\,AU probed in the mid-IR (see Fig.~\\ref{fig:flux_cumule}). This is expected in the case of dust settling (see for instance Fig.~7 in \\D04). The decreasing SED of IM~Lupi in the mid-infrared is very reminiscent of the models including dust settling of \\D04. Thus, we tried to fit the SED without taking into account the mid-infrared fluxes (between 5 and 35\\,$\\mu$m), and hence not considering the silicate features. The resulting Bayesian probabilities still exclude models without dust settling, which do not manage to reproduce simultaneously both the millimetre and far-infrared fluxes. We conclude that dust settling is very likely to occur even in the outer regions of the disc. \\D04 predicts that settled discs may become undetectable in scattered light due to the formation of a self-shadowed opacity structure in the outer disc. This is clearly not the case for IM~Lupi. As noted by \\D04, the self-shadowed structure only appears for low values of the turbulence. This may indicate that the turbulence level in IM~Lupi is large enough to prevent the disc from self-shadowing or that we are observing the outer disc at a relatively early stage of the dust settling process. \\subsection{Evolutionary state of IM~Lupi \\& comparison with other classical T~Tauri stars} Our modelling has lead to a very detailed picture of the disc surrounding IM Lupi. We have already seen that the disc structure of IM Lupi is similar to those of other classical T~Tauri stars. In this section, we compare the signature of dust evolution in the disc with the results obtained for other T~Tauri discs, in order to determine whether it is a singular object or whether it can be considered as representative of other T~Tauri stars. Our conclusions on the degree of dust settling in IM~Lupi result from the strong silicate emission feature and shallow millimetre spectral slope. Other objects show similar characteristics. Thus, Fig.~\\ref{fig:beta} presents the strength of the 10\\,$\\mu$m silicate as a function of the exponent of the opacity law in the millimetre for a set of T~Tauri stars (Table~\\ref{tab:beta}). IM~Lupi is located in the bulk of T~Tauri stars and our results can probably be extrapolated to other sources: detailed analyses of the individual sources are required, but it is likely that at least some of them are undergoing some dust settling. A semblable conclusion was reached by \\cite{Furlan05,Furlan06} and \\cite{D'Alessio06} based on similar evidence \\emph{i.e.}, the presence in the SEDs of 10 and 18$\\,\\mu$m emission silicate bands and the slope and fluxes at \\mbox{(sub-)millimetre} wavelengths. \\begin{figure} \\includegraphics[width=\\hsize]{beta_s10.eps} \\caption{Strength of the 10$\\,\\mu$m silicate features as a function of the millimetre opacity slope for the T~Tauri stars listed in Table~\\ref{tab:beta}. IM~Lupi is represented by the red star. The effect of grain growth and dust settling are schematised by arrows.\\label{fig:beta}} \\end{figure} \\begin{table} \\caption{Millimetre opacity slopes and strengths of the 10$\\,\\mu$m silicate features of T~Tauri stars. References: (1) = \\cite{Lommen07}, (2) = \\cite{Kessler-Silacci06}, (3) = \\cite{Przygodda03}, (4) = \\cite{Rodmann06}, (5) = \\cite{Furlan06}. The source T~Cha presented in \\cite{Lommen07} was not selected, because of the presence of PAH emission. Both \\cite{Rodmann06} and \\cite{Lommen07} take a contribution of optically thick emission into account in their derivation of $\\beta_\\mathrm{mm}$. \\label{tab:beta}} \\centering \\begin{tabular}{lccc} \\hline \\hline Name & $\\beta_\\mathrm{mm}$ & $S_\\mathrm{Peak}(10\\,\\mu\\mathrm{m})$ & Ref. \\\\ \\hline IM Lup & 0.8 $\\pm$ 0.25 & 1.60 & this work \\\\ \\hline HT Lup & 0.4 $\\pm$ 0.5\t& 1.39 & (1, 2)\\\\ GW Lup\t & 0.5 $\\pm$\t 0.5\t& 1.48 & (1, 2)\\\\ CR Cha & 1.5 $\\pm$\t 0.6 & 2.52 & (1, 3)\\\\ WW Cha & 0.8 $\\pm$\t 0.8 & 1.92 & (1, 3)\\\\ RU Lup & 0.8 $\\pm$\t 0.5 & 1.54 & (1, 2)\\\\ RY Tau & 0.8 $\\pm$\t 0.1 & 2.75 & (4, 5)\\\\ FT Tau & 0.9 $\\pm$\t 0.3 & 1.74 & (4, 5)\\\\ DG Tau & 0.7 $\\pm$\t 0.1 & 1.52 & (4, 5)\\\\ UZ Tau E & 0.8 $\\pm$\t 0.1 & 1.65 & (4, 5)\\\\ DL Tau & 1.0 $\\pm$\t 0.2 & 1.1 & (4, 5)\\\\ CI Tau & 1.3 $\\pm$\t 0.4 & 1.5 & (4, 5)\\\\ DO Tau & 0.5 $\\pm$\t 0.1 & 1.19 & (4, 5)\\\\ GM Aur & 1.6 $\\pm$\t 0.2 & 2.54 & (4, 5)\\\\ \\hline \\end{tabular} \\end{table} The tentative correlation between the strength of the silicate feature and millimetre spectral index suggested by \\cite{Lommen07} does not appear as clear in our larger sample of T~Tauri stars. This correlation was interpreted as an indication of fast grain growth in both central and outer regions of the disc. Indeed, as grains grow from sub-micron sizes to several microns, the $10\\,\\mu$m feature becomes weaker and less peaked. When they reach millimetre to centimetre sizes, the slope of the millimetre emission becomes shallower. But, as we have shown, the strength of the silicate features is strongly related to the degree of dust settling. The stronger the settling, the smaller the apparent (\\emph{i.e.} probed in the infrared) grain size, making it difficult to obtain a precise estimate of the actual grain sizes present in the midplane. The detailed analysis of the silicate features indicates a small degree of crystallisation ($< 10\\,\\%$) in spite of a high mass fraction of micrometric (hence evolved) grains. As shown by \\citet[see their Fig.~9 for instance]{Schegerer06}, this is the case for several objects and IM~Lupi is not unique on that aspect. \\cite{Schegerer06} did not find any correlation between the degree of crystallisation and amount of micron-sized grains. \\cite{Furlan05,Furlan06} and \\cite{D'Alessio06} have estimated the degree of dust settling based on the colours and SEDs of a large population of Classical T Tauri stars in the Taurus molecular cloud. They claim that to make a synthetic SED consistent with the median SED of class II objects in Taurus, the dust to gas mass ratio in the disc atmosphere should be around 1\\,\\% (between 0.1 and 10\\,\\%) of the ISM ratio. In our modelling, the grain size distribution and the dust to gas ratio are continuous functions of the height above the midplane. With $\\xi = 0.05$, the dust to gas ratio starts at 1.5 times the ISM value in the disc midplane and reaches 10\\,\\%, 1\\,\\% and 0.1\\,\\% of the ISM value at 2, 3 and 6 scale heights, respectively. These regions roughly correspond to what is defined as the disc surface in \\cite{Furlan05,Furlan06} and \\cite{D'Alessio06}. Although not directly comparable, our estimation of the degree of dust settling appears consistent with the calculations of these authors." }, "0808/0808.3040_arXiv.txt": { "abstract": "Here I discuss possible relations between free precession of neutron stars, Tkachenko waves inside them and glitches. I note that the proposed precession period of the isolated neutron star RX J0720.4-3125 (Haberl et al. 2006) is consistent with the period of Tkachenko waves for the spin period $8.4$~s. Based on a possible observation of a glitch in RX J0720.4-3125 (van Kerkwijk et al. 2007), I propose a simple model, in which long period precession is powered by Tkachenko waves generated by a glitch. The period of free precession, determined by a NS oblateness, should be equal to the standing Tkachenko wave period for effective energy transfer from the standing wave to the precession motion. A similar scenario can be applicable also in the case of the PSR B1828-11. ", "introduction": "Isolated neutron stars (NSs) being non-spherical bodies are expected to demonstrate free precession (for a brief review see, for example, \\citealt{l2003}). However, examples of this phenomena are less than few, and even in rare cases when a precession-like behavior is observed different interpretations can be discussed (even not related to precession, see for example \\cite{rg2006}). The problem of long period free precession in NSs is a long standing one. A NS can precess if it is non-spherical and rotation axis does not coincide with a principal axis. Typically, biaxial objects are discussed, so deviation from spherical symmetry can be described by one parameter -- oblateness (see \\cite{alw2006} for a discussion of triaxial model). Expected values of NS oblateness (due to rotation or influence of strong magnetic fields) can naturally lead to precession periods about one year. The precession period is equal to $P_\\mathrm{prec}=P/\\epsilon$. Here $P$ is the spin period of a NS, $\\epsilon$ -- its oblateness, and $P_\\mathrm{pres}$ -- precession period. Measured precession periods require oblateness about $10^{-8}$. However, discussing dynamics of NSs it is necessary to take into accout the network of superfluid vortices inside them. The neutron superfluid liquid in the interior of a NS participates in rotation via formation of quantized vortex lines. The density of these lines per unit area is $n=2\\Omega / k$. Here $\\Omega=2\\pi /P$ is spin frequency, and $k=h/2m_\\mathrm{n}$, where $m_\\mathrm{n}$ is a neutron mass (see, for example, \\citealt{st1983}, Ch. 10). The vortices exist in the core of a NS, where they can interact with superfluid (superconducting) protons and normal electrons, and in the crust, where they can pin to it. Coupling of superfluid neutron vortices with electrons in a core results in damping of free precession \\citep{ao1987}. But the time scale of this damping is long enough, according to these authors. For spin period about 1 second it is $\\sim 400$~--~$10^4$~$P_\\mathrm{prec}$ \\citep{ao1987}. Still, this time scale is much shorter than a NS age, so some excitation mechanism is necessary for precession. As it is discussed below, in the presented model excitation is due to a glitch. A kind of pinning (``immobilization'') of vortices can also happen in the core due to interactions with magnetic flux tubes (see discussion, for example, in \\citealt{l2007}). In this case, the moment of inertia of ``pinned'' neutrons (which is about $I$) is about 10 times larger, than the moment of the remaining parts of a NS, $I_\\mathrm{c}$. So, $P_\\mathrm{prec}\\sim 0.1 P$. A different kind of problem appears if pinning in the crust is taken into account. For absolute pinning no long period precession is possible. Instead, the period of precession becomes equal to $P (I/I_\\mathrm{p})$, here $I$ is NS moment of inertia, $I_\\mathrm{p}$ is the moment of inertia of pinned superfluid in the crust. Typically it is expected that $I_\\mathrm{p}/I \\sim 10^{-2}$ \\citep{s1977}, and the precession period is just $\\sim100$~$P$ if the absolute pinning is valid. However, \\cite{ao1987} showed that this is not the case due to finite temperature. Because of thermal effect always there is vortex creep which allow the pinned superfluid to follow precession. The best example of a NS with precession-like behavior is PSR B1828-11. The proposed period is about 511 days with a harmonic at 256 days \\citep{sls2000}. Most of discussions related to free precession deal with this source. In particular, the problem of non-existence of long period precession for strong pinning is typically confronted with observaions of PSR B1828-11. Recently, appeared another possible example of long period free precession in NSs. The existence of $\\sim 7$ years precession period in one of a small group of isolated NSs (called XDINS -- X-ray Dim Isolated NSs, or ICoNS -- Isolated Cooling NSs, or the {\\it Magnificent Seven}) -- RX J0720.4-3125 -- was suspected \\citep{hetal2006}\\footnote{A slightly different period $\\sim$ 4.3 years was proposed by van Kerkwijk and Kaplan (2007) based on timing analysis. However, these authors consider a model with a glitch to fit better due to a rapid change in spectral properties, see van Kerkwijk et al. (2007).}. So, this object was added to the list, and the paradoxical situation of long precession in presence of superfluid vortices was reconsidered by \\cite{l2006,l2007}. This author proposed that either protons in NS interiors are type I superconductors, or neutrons in the outer core are normal (i.e., not superfluid). More recently \\cite{gaj2007} demonstrated that for long spin period and small precession angles NSs can have long precession periods (note, that for PSR B1828-11 the precession angle is proposed to be small, about few degrees \\citep{sls2000}, but for RX J0720.4-3125 it can be larger, $>$~10 degrees \\citep{hetal2006}). So, according to \\cite{gaj2007}, the conclusion by \\cite{l2006} and other authors that in the strong drag regime $P_\\mathrm{prec}\\sim 0.1 P$ can be under doubt due to a short wavelength instability. Clearly, the problem of free precession in NSs is far from being solved completely. In this brief note, based on coincidence between Tkachenko wave period and precession period in cases of PSR B1828-11 and RX J0720.4-3125, I discuss a mechanism to support precession in isolated NSs. ", "conclusions": "Absence of free precession in absolute majority of isolated NSs indicates that this phenomena needs some rare coincidence in properties of a NS. Here it is proposed that it is necessary to have: \\begin{itemize} \\item $P_\\mathrm{T}\\approx P_\\mathrm{prec}$, \\item a glitch to generate Tkachenko waves. \\end{itemize} Instead of the proposal by \\cite{l2007} -- {\\it \"A slowly-precessing neutron star cannot glitch\"} -- I propose another: {\\it slow precession is powered by glitches via Tkachenko waves}. Observations of RX J0720.4-3125 are roughly consistent with this scenario. On the other hand, in the case of PSR B1828-11 no glitches have been observed. However, it is necessary to study for how long precession can survive after a glitch. If an old estimate by \\cite{ao1987}, $400$~--~$10^4\\, P_\\mathrm{prec}$ is valid, then this time is long enough. If precession is periodically excited by glitches via Tkachenko waves even damping on a time scale of few precession cycles \\citep{l2006} would not contradict observations of RX J0720.4-3125. The glitch in RX J0720.4-3125 reported by \\cite{vketal2007} by its consequencies is similar to the one observed in an anomalous X-ray pulsar (AXP) CXOU J164710.2-455216 \\citep{ietal2007,metal2007}. After the glitch the luminosity of the source was increased, and its spectrum changed. So, the jump in properties of the spectrum and luminosity of RX J0720.4-3125 proposed by \\cite{vketal2007} can be directly related to a glitch, which is weaker than in the case of CXOU J164710.2-455216 (still similarities in behavior of these sources can be considered as a kind of support to the hypothesis of a link between AXPs and XDINS). But the evolution of the NS parameters after the ``jump'' requires precession. Note, that the timing solution before MJD 52821, when a possible \"glitch\" happened according to \\cite{vketal2007}, can be relatively well described by the so-called cubic solution (the second derivative of $\\nu$ is non-zero), see \\cite{vkk2007}. Spectral changes before this date are not very large \\citep{hetal2006, vkk2007}. After MJD 52821 the timing solution is well described by a periodic function, see \\cite{vkk2007} (these authors studied several models with and without a glitch in their two papers), and spectral changes follow this law, too \\citep{hetal2006}\\footnote{However, van Kerkwijk and Kaplan (2007) propose that the period is not close to $\\sim$7 years, but is $\\sim$4.3 years. Still, on a short time scale -- since MJD 52821 -- this is not very certain.}. Based on that, I suggest that the timing residuals might be also explained by a model without (or with small) precession before the glitch, and strong precession after. However, this particular model has never been tested quantitatively against observational data. Glitches naturally can produce thermal afterglows \\citep{hetal1997}. About $10^{38}$~--~$10^{43}$~ergs can be released in a glitch (in the case of RX J0720.4-3125 according to estimates of the increase in spin frequency by \\citealt{vketal2007} this value is closer to $10^{38}$~erg). However, Hirano et al. showed that a thermal response of a NS to a glitch cannot produce a smooth temperature increase on the time scale of years. If surface temperature is increased just by few percent, as it is required by \\cite{vketal2007}, then the brightening lasts just for few days (this corresponds to weak energy release). If we require a temperature rise for a long time, then the effect is too strong \\citep{hetal1997}. So, I conclude that spectral changes on a long time scale should be attributed to precession of the NS. Glitches of AXPs (and soft gamma-repeaters) can be different in nature with respect to radio pulsar glitches, as the former can be related to crust fracture due to superstrong magnetic field. Still, the origin of a glitch is not important for our discussion here. ``Normal'' glitches are quite common for long period pulsars, for example, PSR J1814-1744 with spin period about 4 seconds demonstrated a glitch \\citep{js2006}. So, RX J0720.4-3125 can glitch not only via the mechanism operating in magnetars, but also due to convenient mechanisms proposed for normal radio pulsars. For them one can estimate the reccurence time following \\cite{ab1994}. If the glitch of RX J0720.4-3125 is due to unpinning, then using standard formulae \\citep{ab1994} one obtains that the reccurence time between two succesive glitches is: \\begin{equation} t_{\\mathrm{g}}=\\frac{\\delta \\Omega}{\\dot \\Omega}. \\end{equation} Parameter $\\delta \\Omega$ is the critical value of the difference between the rotation frequencies of normal matter and the superfluid at a boundary layer. $\\delta \\Omega$ itself can be estimated as (Alpar \\& Baykal 1994): \\begin{equation} \\delta \\Omega = \\Delta \\Omega \\frac{I}{I_{\\mathrm{p}}}, \\end{equation} here $I_{\\mathrm{p}}$ is the effective moment of inertia of the region of a pinning layer. Combining these two formulae one obtains the relation for the time between glitches: \\begin{equation} t_{\\mathrm{g}} = \\frac{2\\, I}{I_{\\mathrm{p}}} \\frac{\\Delta \\Omega}{\\Omega} t \\propto t, \\label{eq2} \\end{equation} where $t$ is the age of a pulsar, $t=\\Omega/2\\, \\dot \\Omega$. Surprisingly, the time is about 10 years for RX J0720.4-3125. I.e. it is quite probable to observe one since the discovery of this object. Then, one can expect to see similar phenomenae in other Magnificent seven objects. However, they are less studied, and may be some glitches are missed. Still, to produce luminosity and spectral changes, and to generate Tkachenko waves, it is probably more natural to have a glitch due to a starquake. In the case of the quake model \\citep{ab1994} the time between glitches is longer, about 300 years for RX J0720.4-3125. \\begin{equation} t_{\\mathrm{g}} = \\frac{2(A+B)\\phi \\Delta \\Omega / \\Omega} {I_0\\Omega \\dot \\Omega}. \\end{equation} The estimate above was obtained assuming standard values \\citep{ab1994} $A=10^{52}$ erg, $\\phi=10^{-3}$, $B=10^{48}$~erg, and $I\\sim 10^{45}$~g~cm$^2$ Then, we can be just lucky to find a glitch in $\\sim 10$ years of observations (but note, that it is not the only XDINS observed). Or, in XDINS quakes do not follow the formula for radio pulsars. \\cite{vketal2007} proposed that an accretion episode can be responsible for spectral changes after a ``glitch'' in RX J0720.4-3125. I think that this is not a very probable reason for the origin of the glitch and corresponding changes. It is hardly possible to imagine that if we observe such an episode just after 10 years of observations, other episodes were not frequent during the evolution of this source. With frequent episodes of accretion of light elements a NS should follow a slightly different cooling track \\citep{ketal2006}. Such stars with accreted envelopes are hotter in their youth, but colder after they mature. This is more similar to the properties of 1E1207.4-5209 and Kes 79 \\citep{gh2007}. A remarkable difference between some NSs in supernova remnants (so-called CCOs) and XDINS in the solar vicinity can be due to the existence of accreted envelopes in the former. Note, that we do not observe descendants of 1E1207-like sources in our proximity. If they follow a standard cooling curve, like RX J0720.4-3125 and other XDINS, then they have to be observed. On the other hand, we do not see ancestors of XDINS in supernova remnants. This can be related to the fact that CCOs descendants are too cold at the age of XDINS to be easily detected by ROSAT, and vice versa ancestors of XDINS are not hot enough to be easily found in some supernova remnants. In this note I neglect (as most, if not all, other authors who studied Tkachenko waves in NSs) the influence of interaction between neutron vortex lines and magnetic flux tubes. This interaction can significantly affect the velocity of waves, and so their period, and to damp them.\\footnote{This was noted to us by Prof. M. Ruderman.} The velocity of a Tkachenko wave can be estimated as: \\begin{equation} V_\\mathrm{T}= 1/2 (h \\Omega/2\\pi m_\\mathrm{pair})^{1/2}= (\\Omega k/8 \\pi)^{1/2} \\sim b/P. \\end{equation} Here $m_\\mathrm{pair}=2 m_\\mathrm{n}$, and $b$ is the distance between vortex lines. So, $P_\\mathrm{T}\\sim (R/b) P$ in the simple case when there are no interactions with flux tubes or other complications. When vortices are able to ``communicate'' with the help of numerous magnetic flux tubes, the velocity of the wave can be larger, so the period of Tkachenko-like wave would be shorter. This question should be explored. To conclude, in this short note I proposed that long period precession in RX J0720.4-3125 can be related to Tkachenko waves, generated in a recent glitch. In the proposed model before a glitch precession could be negligible. The critical condition for the free precession excitation is the equality between Tkachenko wave period and the period of free precession." }, "0808/0808.1045_arXiv.txt": { "abstract": "We discuss the main results that were recently published by the Auger Collaboration and their impact on our knowledge of the ultra high energy cosmic rays and neutrinos. ", "introduction": "We have discussed the problems related to the ultra high energy cosmic rays (UHECR) quite many times in recent years after the results of the high resolution Fly's Eye (HiRes) started trickling out several years ago. The cosmic ray energy spectrum extracted from the the HiRes observations indicated that the spectrum suffers a strong decrease at energies above 5$\\times$10$^{19}$ eV~\\cite{HiRes_sp}, consistent with the GZK cutoff~\\cite{GZK} that is due to UHECR interactions with the microwave background. There was thus a strong disagreement with the energy spectrum derived from the previous largest exposure cosmic ray air shower array, Agasa~\\cite{Agasa} that seemed to continue with the same slope above 10$^{20}$ eV. HiRes and Agasa measure the cosmic ray events in different ways. HiRes is a fluorescent detector that follows the shower longitudinal development by detecting the fluorescent light emitted by the Nitrogen atoms in the air ionized by the shower charged particles. The primary cosmic ray energy is obtained from integration of the shower longitudinal profile normalized by the average electron energy loss and the fluorescent efficiency of the air. Agasa, on the other hand, is a classical air shower array that records the particle density at the observation level and uses Monte Carlo calculations to relate it to the primary cosmic ray energy. The dependence of this method on the hadronic interaction model used in the simulations is considered stronger in this approach. In 2007 the first results of the Pierre Auger Observatory (PAO) in Argentina were published. This experiment became the hope for better understanding of UHECR several years ago when its construction started. It is a hybrid air shower experiment that applies both detection methods. PAO consists of an air shower array of area 3,000 km$^2$ where the individual detectors are 10 m$^2$ water tanks placed at distances of 1.5 km in a hexagonal pattern (SD). The shower array is surrounded by 24 fluorescent telescopes (FD) located in four groups of six. The idea is that during cloudless moonless nights when the fluorescent detectors can work the UHECR showers will be observed simultaneously by both types of detectors and this {\\it hybrid} observation will unveil better the shower properties. Here is a brief summary of the Pierre Auger Observatory results. The energy spectrum derived from PAO data~\\cite{auger_sp} shows a feature consistent with the GZK cutoff similar, but not identical to that of HiRes - see Fig.~\\ref{AugerHR}. The spectrum is based on the much higher surface detector statistics normalized to the energy estimates of the fluorescent detector in hybrid events. The problem is that if the energy was estimated only from the SD data (particle density at 1,000 meters from the shower core - $S_{1000}$) the energy scale would be 25\\% higer~\\cite{auger_re}. Formally this is not a big problem because the uncertainty of the energy estimate is 22\\%. Another problem appears when PAO attempts to estimate the type of the primary cosmic ray nuclei. When it is estimated from the depth of the shower maximum measured by the fluorescent detector $X_{max}$ the average primary mass appears to be close to He~\\cite{auger_mu} . If the surface detector density $S_{1000}$ is used to estimate the number of shower muons~\\cite{auger_re} the average primary nuclei appear to be much heavier. The preliminary conclusion from these two disagreements is that showers are absorbed in the atmosphere slower than models predict. The Auger Collaboration succeeded, however, to eliminate a class of UHECR models that predicted that the primary particles are gamma rays. From studies of $X_{max}$ and other shower properties the fraction of gamma rays in the cosmic rays above 10$^{19}$ eV is limited to not more than 2\\%~\\cite{auger_g}. The much smaller statistics at higher energy does not allow for such strong limits, but at least the bulk of UHECR, which we consider extra galactic, consists of nuclei. The Auger result that impressed the most not only the astrophysical community, but also the general public, was the published correlation of their highest energy events with active galactic nuclei~\\cite{Auger1,Auger2}. Twenty of the 27 events of energy above 5.7$\\times$10$^{19}$ eV (57 EeV) come within 3.1$^o$ of the position of AGN at redshift smaller than $\\leq$0.017 (distance of 71 Mpc) while one expects only 7 in case of isotropic source distribution. The 3.1$^o$ circle around the event direction (that was obtained from a scan in energy, angle and distance) is understood as scattering in the galactic magnetic field plus about 1$^o$ Auger angular resolution. Accounting for the number of scans the chance probability is well below 1\\%. If the events that come from galactic latitude less than 10$^o$ (where catalogs are not full and cosmic ray may scatter more) are excluded then 19 out of 21 events are closer than 3.1$^o$ from nearby AGN. Since the Auger collaboration published the arrival directions and energies of these 27 events in Ref.~\\cite{Auger2} this announcement was quickly followed by a number of other analyses and questions, like:\\\\ \\hspace*{5mm} $\\bullet$ why only 71 Mpc when $>$60 EeV cosmic rays should come from distances up to 200 Mpc ?\\\\ \\hspace*{5mm} $\\bullet$ why there are no events from the direction of the Virgo cluster, that contains powerful AGN ?\\\\ \\hspace*{5mm} $\\bullet$ what does the concentration of events from directions close to Cen A means ?\\\\ Many of these questions were also asked by the Auger Collaboration~\\cite{Auger2}. ", "conclusions": "The first results of the Auger Collaboration did improve our knowledge of the ultra high energy cosmic rays. They are fully consistent with a GZK feature in the cosmic ray spectrum as suggested previously by the HiRes Collaboration. Being a hybrid experiment Auger demonstrated that the energy estimate by fluorescent detectors and surface air shower array is different, as previously suggested by the Agasa and HiRes spectra. Now the UHE community is searching for the reasons for this disagreement that may hide in the currently used hadronic interaction models. Auger started the direct search for the sources of UHECR by studies of the correlation of their highest energy events with active galactic nuclei. The published correlation inspired many other different analyses of correlation with powerful astrophysical objects. This work is in progress now and the UHECR sources will be identified eventually when the Auger statistics grows - it will double at the end of August 2008. Since the current UHECR energy spectrum can be fit with several different models it is not obvious that we know better the expected flux of cosmogenic neutrinos. This may only happen when an unique fit of the spectrum is achieved.\\\\[3truemm] {\\bf Acknowledgments} This talk is based on collaboration with many colleagues including Peter Biermann, Daniel DeMarco, Thomas Gaisser, J\\\"{o}rg Rachen, \\& David Seckel. The author was supported in part by NASA APT grant NNG04GK86G." }, "0808/0808.3781_arXiv.txt": { "abstract": "{Nonthermal radiation observed from astrophysical systems containing relativistic jets and shocks, e.g., gamma-ray bursts (GRBs), active galactic nuclei (AGNs), and Galactic microquasar systems usually have power-law emission spectra. Recent PIC simulations of relativistic electron-ion (electron-positron) jets injected into a stationary medium show that particle acceleration occurs within the downstream jet. In the presence of relativistic jets, instabilities such as the Buneman instability, other two-streaming instability, and the Weibel (filamentation) instability create collisionless shocks, which are responsible for particle (electron, positron, and ion) acceleration. The simulation results show that the Weibel instability is responsible for generating and amplifying highly nonuniform, small-scale magnetic fields. These magnetic fields contribute to the electron's transverse deflection behind the jet head. The ``jitter'' radiation from deflected electrons in small-scale magnetic fields has different properties than synchrotron radiation which is calculated in a uniform magnetic field. This jitter radiation, a case of diffusive synchrotron radiation, may be important to understand the complex time evolution and/or spectral structure in gamma-ray bursts, relativistic jets, and supernova remnants. } \\FullConference{Workshop on Blazar Variability across the Electromagnetic Spectrum\\\\ April 22-25 2008\\\\ Palaiseau, France} \\begin{document} ", "introduction": "Shocks are believed to be responsible for prompt emission from gamma-ray bursts (GRBs) and their afterglows, for variable emission from blazars, and for particle acceleration processes in jets from active galactic nuclei (AGN) and supernova remnants (SNRs). The predominant contribution to the observed emission spectra is often assumed to be synchrotron- and inverse Compton radiation from these accelerated particles for gamma-ray bursts [1-7] and for AGN jets [8-13]. It is assumed that turbulent magnetic fields in the shock region lead to Fermi acceleration, producing higher energy particles \\cite{fermi49,blaneich97}. To make progress in understanding emission from these object classes, it is essential to place modeling efforts on a firm physical basis. This requires studies of the microphysics of the shock process in a self-consistent manner \\cite{p05b,wax06}. ", "conclusions": "We have started to calculate emission directly from our simulations using the same method described in the previous section. In order to calculate the (jitter-like) synchrotron radiation from the particles in the electromagnetic fields generated by the filamentation instability, the retarded electric field from a single particle is Fourier-transformed to give the individual particle spectrum as described in the previous section. The individual particle spectra are added together to produce a total spectrum over a particular simulation time span \\cite{hedeT05,hedeN05}. It should be noted that for this calculation very large simulations over a long time ($t_{\\rm s}$) are required using a small time step ($\\Delta t$) in order to increase the upper frequency limit to the spectrum (Nyquist frequency $\\omega_{\\rm N} = 1/2\\Delta t$). Frequency resolution is limited by the time span ($\\Delta \\omega = 1/t_{\\rm s}$) \\cite{hedeT05,hedeN05}. For a case with the time step $\\Delta t = 0.01/\\omega_{\\rm pe}$ and the time span $t_{\\rm s} = 50/\\omega_{\\rm pe}$, a calculated spectrum will have the highest frequency, $50\\omega_{\\rm pe}$ and the frequency resolution (the lowest frequency), $0.02\\omega_{\\rm pe}$. $\\omega_{\\rm pe}$ is calculated with an appropriate plasma density. Simulations over a long time allow us to obtain multiple spectra at sequential time spans so the spectral evolution can be calculated. Synthetic spectra obtained in the way we have described above should be compared with GRB prompt and afterglow observations. In the case of AGN jets, diffusive synchrotron radiation has already been invoked by several works \\cite{flei06b,FT07b,mao07} to reproduce spectra of 3C273, M87 and Cen A knots from radio to X-rays. For TeV blazars, taking into account the relative importance of the energy densities contained in the small-scale and large-scale magnetic fields may be an elegant alternative to the choice of a broken power-law for the energy distribution of radiating electrons." }, "0808/0808.0261_arXiv.txt": { "abstract": "Conditions are obtained for the existence of a warm inflationary attractor in the system of equations describing an inflaton coupled to radiation. These conditions restrict the temperature dependence of the dissipative terms and the size of thermal corrections to the inflaton potential, as well as the gradient of the inflaton potential. When these conditions are met, the evolution approaches a slow-roll limit and only curvature fluctuations survive on super-horizon scales. Formulae are given for the spectral indices of the density perturbations and the tensor/scalar density perturbation amplitude ratio in warm inflation. ", "introduction": "Inflationary models \\cite{guth81,linde82,albrecht82} have proved very sucessful in explaining many of the large scale features of the universe (see e.g. \\cite{Liddle:2000cg}). An essential feature of these inflationary models is their stability, meaning in particular that inflation solutions are attractors in the solution space of the relevant cosmological equations (see e.g. \\cite{Salopek:1990jq,Liddle:1994dx}), at least up until inflation ends and the universe becomes radiation dominated. Without this feature, inflation might never have begun, and certainly would not have lasted long enough to affect the large scale structure of the universe. Warm inflation is an alternative inflationary scenario in which a small but significant amount of radiation survives during the inflationary era due to continuous particle production \\cite{Moss85,bererafang95,berera95}. The coupling between radiation and the inflaton field leads to thermal dissipation and fluctuations in the time evolution of the inflaton field. The stability of the inflationary solutions in warm inflationary models has only been addressed in a limited form previously \\cite{deOliveira:1997jt}, and we will present the full stability analysis here. We shall examine conditions under which warm inflation is an attractor and give conditions for a prolonged period of warm inflation. We shall show that the stability of warm inflation can be related to conditions on two parameters describing the temperature dependence of terms in the inflaton equation of motion which where not taken into account in the earlier stability analysis \\cite{deOliveira:1997jt}. The first condition says that if the dissipation term in the equation of motion falls off too rapidly at low temperature then the temperature is driven to zero, and we fall into the conventional inflationarty scenario. The second condition limits the temperature dependence of the inflaton potential. In many models large dissipation implies large thermal corrections to the potential which prevent warm inflation taking place. This argument is essentially the one first presented in Ref. \\cite{yokoyama99}. Nowadays, we know that there are models where thermal corrections to the inflaton potential are suppressed by supersymmetry, and these models may allow warm inflation as an attractor \\cite{BasteroGil:2006vr,BuenoSanchez:2008nc,review}. The duration of the period of inflation is related to a set of slow-roll paramaters which where introduced in \\cite{Hall:2003zp}. We shall re-derive the slow-roll conditions for warm inflation as part of the stability analysis. A well understood feature of the slow-roll conditons for warm inflation is that they can be less restrictive than the slow-roll condtions for conventional inflation. We stress that we are concerned here with the self-consistency of the warm inflationary scenario for given equations of motion. We shall not address how the equation of motion for the inflaton field is obtained from non-equilibrium thermal field theory. A discussion of the derivation of the equations of motion can be found in a recent review \\cite{review}. However, we would like to point out that some of the critisims of warm inflation have been based on models which do not satisfy the fundamental stability conditions derived here, and are therefore not inconsistent with the validity of warm inflation in general \\cite{Aarts:2007ye}. The stability of warm inflation has consequences for the origin and evolution of cosmological density fluctuations \\cite{Hall:2003zp}. In warm inflation, density fluctuations originate from thermal fluctuations \\cite{bererafang95,berera00}. In particular, the fact that the inflationary solution depends on only one parameter means that only one perturbation mode, the curvature perturbation, survives on super-horizon scales despite the fact that there are entropy perturbations present on sub-horizon scales. We shall give formulae for the spectral indices of the scalar and tensor modes and say a little about the tensor/scalar ratio. ", "conclusions": "We shall recapitulate the main points of this paper. There are conditions on six of the parameters defined in Sect. \\ref{be} for the possibility of a stable period of warm inflation in the early universe: \\begin{itemize} \\item The parameter $Q$ which measures the strength of the friction term must satisfy \\begin{equation} Q>g_*{\\cal P}_S. \\end{equation} where $g_*$ is the effective particle number and ${\\cal P}_S$ is the scalar perturbation power spectrum on large scales. \\item The parameters which describe the inflaton dependence of the effective potential and friction term satisfy \\begin{equation} \\epsilon<<1+Q,\\qquad |\\eta|<<1+Q,\\qquad |\\beta|<<1+Q. \\end{equation} \\item The temperature dependence of the potential and the friction term is restricted by \\begin{equation} |b|<<{Q\\over 1+Q},\\qquad |c|<4. \\end{equation} The condition on $b$ implies that warm inflation is only possible when a mechanism, such as supersymmetry, reduces the size of the thermal corrections to the potential. \\end{itemize} Models of elementary particles exist in which these conditions can be satisfied. The most convincing of these models use a combination of supersymmetry and a two-stage decay process, where there are no direct coupling between the inflaton and the radiation and all thermal effects are supressed by factors of $T/\\Lambda_S$, where $\\Lambda_S$ is the supersymmtry breaking scale \\cite{berera02,BasteroGil:2006vr,BuenoSanchez:2008nc}. When the conditions listed above are satisfied, then the solutions to the equations of motion approach a slow-roll approximation during inflation. As a result, large scale density perturbations have only one degree of freedom, which we identify as the curvature perturbation. (Entropy perturbations can only be introduced by adding extra degrees of freedom to the system.)" }, "0808/0808.1737_arXiv.txt": { "abstract": "We present a new analysis of the long-period variables in the Large Magellanic Cloud (LMC) from the MACHO \\emph{Variable Star Catalog}. Three-quarters of our sample of evolved, variable stars have periodic light curves. We characterize the stars in our sample using the multiple periods found in their frequency spectra. Additionally, we use single-epoch Two Micron All Sky Survey measurements to construct the average infrared light curves for different groups of these stars. Comparison with evolutionary models shows that stars on the red giant branch (RGB) or the early asymptotic giant branch (AGB) often show non-periodic variability, but begin to pulsate with periods on the two shortest period-luminosity sequences (3 \\& 4) when they brighten to $K_s \\approx 13$. The stars on the thermally pulsing AGB are more likely to pulsate with longer periods that lie on the next two P-L sequences (1 \\& 2), including the sequence associated with the Miras in the LMC. The Petersen diagram and its variants show that multi-periodic stars on each pair of these sequences (3 \\& 4, and 1 \\& 2) typically pulsate with periods associated only with that pair. The periods in these multi-periodic stars become longer and stronger as the star evolves. We further constrain the mechanism behind the long secondary periods (LSPs) seen in half of our sample, and find that there is a close match between the luminosity functions of the LSP stars and all of the stars in our sample, and that these star's pulsation amplitudes are relatively wavelength independent. Although this is characteristic of stellar multiplicity, the large number of these variables is problematic for that explanation. ", "introduction": "\\label{intro} Stellar pulsation of giant stars appears to be a ubiquitous and important phenomenon---RR Lyrae and Cepheid variables form the basis for the distance scales we use. Miras and other long-period variables (LPVs) however, are not as well understood, largely because their cool and tenuous atmospheres are dynamic environments with a great diversity of molecular species forming and disassociating as the star pulsates. In recent years however, these stars have attracted increased attention as micro-lensing surveys of the Large and Small Magellanic Clouds (OGLE --- \\cite{1994ApJ...435L.113P}; OGLE II --- \\cite*{1997AcA....47..319U}; MACHO --- \\cite{macho}) have produced large catalogs of LPVs. Well-sampled light curves and excellent photometry give us an opportunity to better understand both the mechanisms behind long-period variables and the physical processes at work in the latest stages of stellar evolution. Before the wealth of data from micro-lensing surveys, LPVs were traditionally classified by the amplitude and stability of their variability in the $V$ band (e.g. The General Catalog of Variable Stars---\\cite{GCVS}). In this scheme, stars with well-defined pulsation are classified as Miras if the amplitude of their variability exceeds 2.5 magnitudes in $V$, and as Semi-Regular Type a (SRa) stars if not. Those with multiple periods, unstable periodicity, or poorly expressed periodicity, are classified as SRb stars. The MACHO Project's survey of the Large Magellanic Cloud (LMC) revealed five parallel sequences of LPVs in period-luminosity space \\citep{cook96}, prompting a classification scheme that uses the period of pulsation as its primary discriminator. \\cite{Wood99} identified the cause of the first three period-luminosity sequences (denoted \\emph{A}, \\emph{B}, and \\emph{C}) as pulsation, and suggested that the two longest period sequences (\\emph{E} and \\emph{D}) could be attributed to binary systems. The stars in Sequence E showed the characteristic light curves of contact binary systems, and Sequence D stars---those with the longest periods---simultaneously exhibited at least one shorter period that was coincident with Sequence B. This is likely to be the LMC equivalent of the ``long secondary periods'' described by \\cite{1963AJ.....68..253H} for Galactic LPVs, although the periods that comprise Sequence D are on average three times shorter than the LSPs listed in \\cite{1963AJ.....68..253H}. \\cite{Wood99} proposed that these stars are composed of accreting binary systems, with the long period caused by partial eclipses due to an unseen, dust-enshrouded companion. In this work the LMC period-luminosity sequences will be named in the manner of \\cite{fraser05}, from shortest to longest period: 4, 3, 2, 1, E, and D. We retain the names D and E from \\cite{Wood99}, but rename his Sequence C to Sequence 1 and count up toward the shorter periods. This approach provides for a graceful way to accommodate additional short-period sequences. Indeed, the use of 2MASS $K_s$ magnitudes as the luminosity indicator caused a split in the original Sequence B---producing Sequences 2 and 3; \\cite{2003MNRAS.343L..79K}. A fifth sequence was identified by \\cite{2004AcA....54..129S} through the examination of all significant frequencies of these stars instead of just the strongest frequency. In general, stars brighten and redden as they evolve along the Red Giant Branch (RGB) and the Asymptotic Giant Branch (AGB). Since the typical $J-K_s$ color of the stars in each sequence reddens as we progress from the short-period Sequence 4 to the longer period Sequence 1, this suggests that evolution proceeds from shorter periods toward longer periods, at least for Sequences 1--4 \\citep{fraser05}. In fact, the low luminosity bases of Sequences 2, 3, and 4 are heavily populated by RGB stars \\citep{2002MNRAS.337L..31I, 2003MNRAS.343L..79K, 2004MNRAS.347L..83K, 2004MNRAS.347..720I}. Above the tip of the RGB, models of AGB stars that include the effects of mass loss confirm that stars continue their evolution to higher luminosities \\citep{1993ApJ...413..641V}. In period-luminosity space, the Mira and SRa stars are not clearly separated, with SRa stars found throughout Sequences 1--4. \\cite{2004AcA....54..129S} identified a more useful division for LPVs in the LMC and Small Magellanic Cloud (SMC) than the Mira/SRa/SRb system. In this system an LPV is classified as an OSARG (Ogle Small Amplitude Red Giant) if one of its three strongest periods falls onto Sequence 4, and as an SRV/Mira if not. This division separates LPVs into two groups with a variety of distinct properties, and it also shows that Sequence D is composed of two different populations: a dimmer population that covers a relatively broad period range, and a more luminous, redder population that shows a tighter period-luminosity relationship (the color changes described here can be seen in \\cite{fraser05}). Stars in Sequence E show ``ellipsoidal'' light curves \\citep{2004AcA....54..347S}, where the brightness modulation is due to the gravitational distortion of one member of a close binary system. These light curves exhibit dual minima of unequal depths, but the effect is small enough that most methods (including ours) find a period for these systems that is half of the orbital period. We refer to this period as the Fourier period and use it to distinguish Sequences E and D in our plots. When stars in Sequence E are plotted on the period-luminosity diagram at their orbital period (as in \\cite{2004AcA....54..347S} and \\cite{2006ApJ...650L..55D}) they smoothly join Sequence D. This, along with a recent analysis of OGLE light curves in \\cite{2007ApJ...660.1486S}, supports the binary companion explanation for Sequence D put forth by \\cite{Wood99}, as does a radial velocity study of several of these stars \\citep{2006MmSAI..77..537A}. In \\cite{fraser05} (hereafter Paper I) we used the MACHO and 2MASS magnitudes and colors to characterize LMC stars in each period-luminosity sequence, and classified the stars in each sequence as Miras, SRa, or SRb. At that time, only 52 percent of the stars in the color- and magnitude-defined sample had a well-determined period in the MACHO Variable Star Catalog. In this work, we expand the successfully analyzed stars to 93 percent of our sample, as well as consider their multi-periodic properties. We also use our results to describe the characteristic variability at each of the stages of RGB and AGB evolution by comparison with population synthesis models, including the stars that show very weak or non-existent periodicity. ", "conclusions": "\\label{discussion} The analysis of the frequency spectra of variable stars has been used with great success for characterizing light curve morphology, identifying binary stars, and constraining observed pulsation modes in studies of Cepheids \\citep{1981ApJ...248..291S}, Beat Cepheids \\citep{1995AJ....109.1653A}, Type II Cepheids and RV Tauri stars \\citep{1998AJ....115.1921A}, RR Lyrae \\citep{2000ApJ...542..257A}, and LPVs (see \\S \\ref{intro}). Here we discuss how these techniques contribute to the understanding of the origin of Sequence D and the relationship between the different LPV stages and stellar evolution. \\subsection{Sequence D} One-third of the stars in our sample exhibit the ``long secondary period'' phenomenon. Although the exact mechanism for Sequence D is still unknown, there are many reasons to think that it is correlated with binary systems. In this paper, the evidence includes the similarity between the luminosity functions of Sequence D and our entire sample (Figure \\ref{LumFunctions}, right panel), and the similar amplitudes of Sequence D's average infrared light curves across the 2MASS bands. Other evidence includes the smooth connection between Sequence D and Sequence E---when E is plotted at its orbital period \\citep{2004AcA....54..347S}, the presence of ellipsoidal light curves \\citep{2007ApJ...660.1486S}, and radial velocity studies of these stars \\citep{2006MmSAI..77..537A}. However, there are also several significant features of Sequence D that are unique, or show differences from Sequence E. Sequence D does not suffer from period halving like Sequence E does, and Sequence E stars often have the third and fourth harmonics present in their frequency spectra, while Sequence D stars show the second harmonic instead (\\S \\ref{multiP}). \\cite{2006ApJ...650L..55D} found that in an amplitude-luminosity plot, stars in Sequence D follow a different pattern than in Sequence E. We see that Sequence E appears to be a continuation of only the small-amplitude ($<0.2$ Blue peak-to-peak) group of stars in Sequence D, those stars that roughly correspond to the OSARGs of \\cite{2004AcA....54..129S}. This is consistent with the comparison of the average light curve amplitudes in Blue and $K_s$, where only the lower amplitude Sequence D stars were a good match to Sequence E. Finally, Sequence D's infrared light curves \\emph{lead} the optical light curves by 10--15 percent, which is a feature unique to this sequence. These facts do not necessarily preclude a binary star mechanism for Sequence D, but they are useful constraints for proposed mechanisms. We note, as an additional constraint, that stars associated with Sequence 1 are far less likely to have a period lying on Sequence D, and that Sequence D is not observed in the LMC below $K_s \\approx 13.7$. The model proposed by \\cite{2007ApJ...660.1486S}, based on the original model proposed by \\cite{Wood99}, is that of a binary system in which the mass lost from the red giant is concentrated near the companion, and regularly obscures the red giant. The wide range of observed light curve amplitudes for Sequence D stars, from 0.1 up to five magnitudes in the MACHO Blue filter \\citep{fraser05}, can be readily explained by the projection effects of different inclination angles in a binary system. If Sequences E and D are truly composed of binary systems, then the population of binary stars in our sample includes stars with either their primary or secondary Fourier period lying within the boundaries of these sequences (e.g. the top panel of Figure \\ref{LightCurves} shows a star whose secondary Fourier period lies on Sequence D). After removing periods identified with the One-Year Artifact, we find that 48 percent of the stars in our sample have variability associated with Sequences E or D (see Table \\ref{sequences}). For comparison, \\cite{1991A&A...248..485D} found a binary fraction of approximately 40 percent for nearby solar-type stars, and \\cite{1997AJ....113.2246R} found a fraction of approximately 35 percent for low-mass stars, a trend in mass which is discussed in \\cite{2006ApJ...640L..63L}. It is reasonable to assume that we cannot see pole-on binaries, and there is no reason to think that all binaries have periods shorter than four years, both of which imply that 48 percent underestimates the total percentage of binaries seen in the LMC. This is a serious problem for any explanation of Sequence D that relies solely on binary systems. However, it is very likely that a subset of these stars do show variability due to binarity, perhaps stars in one of the populations that can be separated by color or amplitude. \\subsection{Comparison to Evolutionary Models} \\begin{figure} \\begin{center} \\includegraphics[width=0.9\\textwidth]{JK} \\caption[Average $J-K_s$ color for each sequence.]{Average $J-K_s$ color for each sequence, as well as stars in the One-Year Artifact and stars in the ``background'' of the Fourier period-luminosity diagram. $J-K_s$ colors are calculated for 0.5 magnitudes bins and shown for bins with more than 40 stars.} \\label{JK} \\end{center} \\end{figure} We can begin to characterize the evolution of LPVs by comparison in color-magnitude space to models. \\cite{2003A&A...403..225M} produced a population synthesis model of the $K_s$ vs. $J-K_s$ color-magnitude diagram (CMD) of the LMC using the RGB and Early AGB (E-AGB) evolutionary models of \\cite{2000A&AS..141..371G}, and a preliminary version of their thermally pulsing AGB star models (published later in \\cite{2007A&A...469..239M}). Their Figure 12 illustrates the luminosities and colors corresponding to major phases in an LMC star's evolution from RGB, through the Early AGB, and finally to the thermally pulsing AGB (including a transformation to carbon-dominated atmospheres for some stars). For comparison, our Figure \\ref{JK} shows the average $J-K_s$ color binned in magnitude for each of the Sequences 1--4, E, D, as well as the One-Year Artifact and the background population of stars. The color of the ``background'' stars match Galactic disk turn-off stars and LMC intermediate-mass stars on the Early AGB in the synthetic CMD of \\cite{2003A&A...403..225M}. As shown in Figure 12, the $J-K_s$ color is more effective at distinguishing evolutionary stages in the AGB than in the RGB. However, with the additional information from the luminosity functions, we can also estimate the importance of RGB stars to each sequence. A distinct peak in the luminosity function at the tip of the RGB ($K_s=12.3$, \\cite{2000ApJ...542..804N}) is widely taken to indicate that the majority of the stars dimmer than this point are themselves on the RGB \\citep{2002MNRAS.337L..31I, 2003MNRAS.343L..79K, 2004MNRAS.347L..83K, 2004MNRAS.347..720I, fraser05}. The stars in our sample that show very weak or nonexistent periodicity (the 24 percent identified with the One-Year Artifact) are predominantly RGB stars. The left two panels of Figure \\ref{LumFunctions} show the luminosity functions of stars in the One-Year Artifact with Sequences 3 and 4. Below the tip of the RGB the population of One-Year Artifact stars dominates, suggesting that very weak or non-periodic variability is common among RGB stars. Above the tip of the RGB there is a much closer correspondence between the One-Year Artifact and Sequences 3 and 4. Thus it appears that stars at the dimmest luminosities in our sample vary aperiodically while on the RGB, but most begin to show periodic behavior when they brighten to $K_s \\approx 13$. After passing off of the tip of the RGB, stars may pulsate with shorter periods and lower luminosity as RR Lyrae on the horizontal branch. They become LPVs again as they ascend the Early AGB (i.e. prior to the first thermal pulse or helium shell flash). \\cite{2004AcA....54..129S} used the slight offset in period between RGB and AGB stars to show that AGB stars pulsate alongside RGB stars below the tip of the RGB. Evolution proceeds to brighter luminosities until the onset of thermal pulses, which begin at $K_s \\approx 12$ on the synthetic CMD from \\cite{2003A&A...403..225M}. Stars primarily populate Sequences 1 and 2 above this luminosity. \\cite{1996A&A...311..509W} predict that thermal pulses will create large modulations in luminosity and the pulsation period (and mode) in AGB stars with timescales of thousands of years. The models of \\cite{2007A&A...469..239M} substantially agree with these predictions, and show large changes in pulsation period due to mode switching as a direct result of a thermal pulse. Period changes in LPVs are well known \\citep{2005AJ....130..776T} but only a small percentage of stars at any one time should be undergoing a thermal pulse due to the short timescales of thermal pulses relative to the long inter-pulse period. The observed period changes of LPVs are not well explained by the effects of thermal pulses alone. The middle two panels of Figure \\ref{LumFunctions} compare the luminosity functions of the numbered sequences. The relative importance of the two giant branches shifts from the RGB to the AGB as we move to Sequences 1 and 2. Additionally, the peak number of stars above the tip of the RGB in each sequence is found at higher luminosity from Sequence 4 to Sequence 1. The OSARG versus SRV/Mira distinction of \\cite{2004AcA....54..129S}, by virtue of its definition, roughly corresponds to a division between two pairs of sequences: Sequences 3 and 4, and Sequences 1 and 2. This division is also clearly seen in the observed period ratios (Figure \\ref{Petersen2}). Considering the synthetic CMD from \\cite{2003A&A...403..225M}, OSARGS are closely related to RGB and E-AGB stars, while the SRV/Mira stars are more closely related to thermally pulsing AGB stars. At $K_s \\approx 11$ the synthetic CMD predicts the formation of carbon stars, and the typical $J-K_s$ colors of each sequence diverge (Figure \\ref{JK}). Only Sequences 1, 2, and D redden to the expected $J-K_s$ color of the carbon star tail of \\cite{2003A&A...403..225M}. Their models also show that these stars do not evolve in brightness after this stage, so the observed range of carbon star luminosities in our sample may be interpreted as a range of stellar masses. \\cite{2007A&A...469..239M} show that only stars between 1.2 and 2.5 $M_\\odot$ undergo a dredge-up that can bring the results of nuclear burning to the atmosphere without quickly destroying it through hot bottom burning. Since Miras exist on Sequence 1 at luminosities both above and below $K_s=11$ \\citep{fraser05}, not all large amplitude pulsators are carbon stars. Also, not all carbon stars have such red $J-K_s$ colors: \\cite{2004A&A...425..595G} found carbon stars on the shorter period sequences---the popular color-cut of $J-K_s \\ge 1.4$ appears more effective at segregating M stars, which are rarely this red. Using $J-K_s>1.4$ to select areas that include only carbon stars, we see that many occupy the highest luminosities of both Sequences 1 and 2, as also seen in \\cite{2007arXiv0710.0953L}. Approximately 40 percent of the carbon stars lie on Sequence 2, similar to the prediction of \\cite{2003A&A...403..225M}, who fit the observed $K_s$ vs. $J-K_s$ CMD by assuming a 50 percent mix of fundamental and first-overtone pulsation among the carbon stars in their model. Figure \\ref{JK} shows that the stars on Sequence 1 evolve to redder $J-K_s$ colors than on Sequence 2, presumably due to increased mass loss driven by fundamental-mode pulsation \\citep{2007A&A...469..239M}. Some stars on the long-period extreme of Sequence 1 are underluminous for their color, which may be self-extinction due to dusty outflows. Beyond this point, stars begin their rapid post-AGB evolution, and they quickly move out of the color-magnitude space of our sample. \\subsection{Brief Summary of LPV Evolution} LMC stars on the RGB and AGB are characterized by the presence of multiple long periods that show increasing length and amplitude as these stars evolve. Apart from the presence of the long secondary period phenomenon, which appears in approximately half of the stars in our sample, stars initially vary non-periodically, and only later begin to pulsate with periods of 20--120 days on Sequences 3 and 4. The Petersen diagram and its variants show that stars with their primary period on either one of these sequences often have their secondary period on the other sequence. Furthermore, the period changes between the first half and last half of the MACHO light curves suggest that the amplitudes of these pairs of periods are very similar for stars on the inside edges of these two sequences. Similar results are obtained for stars which have evolved to the luminosity at which thermal pulses begin on the AGB ($K_s \\approx 12$); these stars are usually found on Sequences 1 and 2 with periods of 45--500 days. This supports the arguments of \\cite{1986MNRAS.219..525W} and \\cite{agbstars2}, that a star excites different pulsation modes in turn as it evolves. At the points where the dominant modes switch, we observe pulsation in multiple periods with near equal strengths. The increase in the luminosity of the maximum of the luminosity function above the tip of the RGB also lends support to this argument. The highest luminosity stars on Sequences 1 and 2 have become carbon stars. After this point, the mass-loss rate of these stars increases drastically and they rapidly evolve out of our sample of luminous red stars." }, "0808/0808.0507_arXiv.txt": { "abstract": "We present the results of a recent reverberation-mapping campaign undertaken to improve measurements of the radius of the broad line region and the central black hole mass of the quasar PG\\,2130+099. Cross correlation of the 5100\\,\\AA\\ continuum and H$\\beta$ emission-line light curves yields a time lag of 22.9$^{+4.4}_{-4.3}$ days, corresponding to a central black hole mass $M_{\\rm BH} = $ ($3.8 \\pm 1.5 $)$ \\times 10^{7} M_{\\odot}$. This value supports the notion that previous measurements yielded an incorrect lag. We re-analyzed previous datasets to investigate the possible sources of the discrepancy and conclude that previous measurement errors were apparently caused by a combination of undersampling of the light curves and long-term secular changes in the H$\\beta$ emission-line equivalent width. With our new measurements, PG\\,2130+099 is no longer an outlier in either the $R_{\\rm BLR}$--$L$ or the $M_{\\rm BH}$--$\\sigma_*$ relationships. ", "introduction": "Reverberation mapping uses observations of continuum and emission-line variability to probe the structure of the broad line region (BLR) in active galactic nuclei (Blandford \\& McKee 1982; Peterson 1993). It has been extensively used to estimate the physical size of the BLR and the mass of central black holes in active galactic nuclei (AGN). The observed continuum variability precedes the observed emission-line variability by a time related to the light travel time across the BLR; by obtaining an estimate of the time delay, or ``lag'' $\\tau$, between the change in continuum flux and the change in emission-line flux, one can estimate the size of the BLR. While this is an extremely effective method, high-quality datasets are difficult to obtain, as they require well-spaced observations over long timescales. To date, over three dozen AGN have estimated black hole masses obtained using reverberation methods. In addition to being important physical parameters, the BLR radius and central black hole mass ($M_{\\rm BH}$) measurements are crucial in calibrating relationships between different properties of AGN. A useful relationship that has emerged is the correlation between the radius of the BLR ($R_{\\rm BLR}$) and the optical luminosity of the AGN (e.g. Kaspi et al.\\ 2000, 2005; Bentz et al.\\ 2006). Another key relationship is that between $M_{\\rm BH}$ and bulge stellar velocity dispersion ($\\sigma_*$) which is seen in both quiescent (Ferrarese \\& Merritt 2000; Gebhardt et al.\\ 2000a; Tremaine et al.\\ 2002) and active galaxies (Gebhardt et al.\\ 2000b; Ferraresse et al.\\ 2001; Nelson et al.\\ 2004; Onken et al.\\ 2004; Dasyra et al.\\ 2007). These two relationships are critical because they allow us to estimate the masses of black holes in AGN from a single spectrum (see McGill et al.\\ 2008 for a recent summary) --- we estimate $R_{\\rm BLR}$ from the AGN luminosity, the velocity dispersion of the BLR is determined by the emission-line width, and the quiescient $M_{\\rm BH}$--$\\sigma_{*}$ relationship provides the calibration of the reverberation-based mass scale. Obtaining high-quality single-epoch spectra is much less observationally demanding than obtaining high-quality reverberation measurements; once relations such as these are properly calibrated, we can measure $M_{\\rm BH}$ in many more objects than would otherwise be possible. An extensive set of objects with reliable black hole masses allows us to explore the connection between black holes and AGN evolution over cosmologically interesting timescales. PG\\,2130+099 has been a source of curiosity because it is an outlier in both the $M_{\\rm BH}$--$\\sigma_{*}$ and $R_{\\rm BLR}$--$L$ relations. Previous measurements obtained by Kaspi et al.\\ (2000) and reanalyzed by Peterson et al.\\ (2004) found H$\\beta$ lags of about 180 days. These measurements yield a black hole mass $M_{\\rm BH}$ upwards of $10^{8}$ $M_{\\odot}$. At luminosity $\\lambda L_{\\lambda}$(5100\\,\\AA)=($2.24\\pm0.27$)$\\times10^{44}$ erg s$^{-1}$, this places PG\\,2130+099 well above the $R_{\\rm BLR}$--$L$ relationship (Bentz et al.\\ 2006). Dasyra et al.\\ (2007) also note that PG\\,2130+099 falls above the $M_{\\rm BH}$--$\\sigma_{*}$ relationship. Together these suggest that this discrepancy could be caused by measurement errors in the BLR radius. Our suspicions are also fueled by two other factors. First, the optical spectrum of PG\\,2130+099 is similar to that of narrow-line Seyfert 1 (NLS1) galaxies, which are widely supposed to be AGN with high accretion rates relative to the Eddington rate (see Komossa 2008 for a recent review). However, the accretion rate derived from this mass and luminosity (under common assumptions for all reverberation-mapped AGN as described by Collin et al.\\ 2006) is quite low compared to NLS1s, again suggesting that $M_{\\rm BH}$ and therefore $\\tau$ are overestimated. Second, as we discuss further below, a lag of approximately half a year on an equatorial source means that fine-scale structure in the two light curves will not match up in detail, and cross-correlation becomes very sensitive to long-term secular variations that may or may not be an actual reverberation signal. For these reasons, we decided to undertake a new reverberation campaign to remeasure the H$\\beta$ lag for PG\\,2130+099. In this paper, we present a new lag determination from this campaign. We also present a re-analysis of the earlier dataset that suggests that the true lag is consistent with our new lag value and investigate possible sources of error in the previous analysis. ", "conclusions": "Analysis of previous datasets of PG\\,2130+099 yields lags greater than 150 days and $M_{\\rm BH}$ values above $10^{8}$ $M_{\\odot}$, in both cases approximately an order of magnitude larger than our new value. To investigate the source of these discrepancies, we closely scrutinized the previous dataset, which consists of data obtained at Steward Observatory and Wise Observatory (Kaspi et al.\\ 2000). The 5100\\,\\AA\\ continuum and emission-line light curves, along with their respective CCF and ZDCFs, are shown in Figure \\ref{fig:f4}. We first ran the full light curve through the time series analysis as described in section 2.3 to confirm the previous results, and successfully reproduced a lag measurement of $\\sim168$ days, in agreement with Kaspi et al.\\ (2000) and Peterson et al.\\ (2004). However, visual inspection of the light curves suggests that these values were in error; we were able to identify several features that are present in the continuum, H$\\beta$ and H$\\alpha$ light curves that are quite visibly lagging on timescales shorter than 50 days, as we show below. It is clear that the Kaspi et al.\\ results are dominated by the data spanning the years 1993-1995, as this period contains the most significant variability as well as time sampling that is sufficient to resolve features in the light curves. We show the light curves for all spectral features measured by Kaspi et al.\\ during this time period in Figure \\ref{fig:f5}. Over this period, we can match behavior in the continuum with the behavior of the emission lines, most noticeably in the data from 1995. The most prominent features in the light curves are the maximum in the continuum at the end of the 1994 series and the maxima in the lines at the beginning of the 1995 series, which, based on the continuum variations, one would expect to be lower relative to the fluxes in 1994. However, inspection of the relative flux levels of the lines and continuum in the 1994 and 1995 data reveals that the equivalent width of the emission lines changes during this timespan, which results in a rise in emission-line flux apparently unrelated to reverberation. The maximum in the continuum and those in the emission lines do not correspond to the same features--- the maxima in the emission line light curves in 1995 correspond to the similar feature in the continuum at the beginning of 1995. Because there are no observations during the six months or so between these features, the CCF and ZDCF lock onto the two unrelated maxima that are separated by 196 days and yield lags of $\\sim$200 days over this three-year span. The emission lines likely continued to increase in flux during the time period for which there were no observations. Inspection of the light curve segments in Figure \\ref{fig:f5} reveals similar structures in each of them within a given year. To quantify the time delays between the continuum and lines on short timescales, we cross-correlated each individual year, this time using only flux randomization in the Monte Carlo realizations, as there were too few points in each year to use random subset sampling, so the uncertainties are underestimated. The resulting lags are given in Table \\ref{Table:lags}. Again we must consider that the CCF cannot produce a lag that is greater than the duration of observations, so the individual CCFs for these three years are limited to short delays. However, the presence of the 1995 feature in both the continuum and emission-line light curves is unmistakable and it is extremely unlikely that it lags by a value exceeding the sensitivity limit of the CCF. From this we surmise that the discrepancies in measured lag values are mostly a result of large time gaps in the data and/or underestimation of the error bars in the Kaspi et al.\\ data. Welsh (1999), and even earlier, P\\'{e}rez, Robinson, \\& de la Fuente (1992), pointed out that emission-line lags can be severely underestimated with light curves that are too short in duration, particularly if the BLR is extended: certainly our 98-day campaign is insensitive to lags as long as $\\sim180$\\,days. Is it possible that the original lag determination of Kaspi et al.\\ (2000) and Peterson et al.\\ (2004) is correct (or more nearly so) and we have been fooled by reliance on light curves that are too short? We think not, based on (1) the reasonable match between details in the continuum and emission-line light curves for four different observing seasons (1993, 1994, 1995, and 2007), (2) the improved agreement with the $R_{\\rm BLR}$-$L$ relationship with the smaller lag (demonstrated in Figure \\ref{fig:f6}), and (3) the improved agreement with the $M_{\\rm BH}$-$\\sigma_*$ relationship with the smaller lag (Figure \\ref{fig:f7}). It is also worth pointing out the difficulty of accurately measuring a $\\sim 180$\\,day lag, particularly in the case of equatorial sources which have a relatively short observing season. The short observing season, typically 6-7 months, means that there are very few emission-line observations that can be matched directly with continuum points: the observed emission-line fluxes represent a response to continuum variations that occurred when the AGN was too close to the Sun to observe. Welsh (1999) has also pointed out the value in ``detrending'' the light curves--- removing long-term trends by fitting the light curves with a low-order function can reduce the bias toward underestimating lags. In this particular case, we find that detrending has almost no effect: in particular, the highest points in the continuum (in late 1994) and the highest points in the line (early in the 1995 observing season) remain so after detrending, and both the interpolation CCF and ZDCF weight these points heavily. It is also interesting to note that the peak in the ZDCF agrees with the peak of the interpolation CCF (Figure \\ref{fig:f4}), which demonstrates that the $\\sim180$\\,day lag is {\\em not} simply ascribable to interpolation across the gap between the 1994 and 1995 observing seasons. The data from our 2007 campaign are not ideal: the amplitude of variability was low, the time sampling was adequate, but only barely, and the duration of the campaign was short enough that we lack sensitivity to lags of $\\sim 50$\\,days or longer. But the preponderance of evidence at this point argues that our smaller lag measurement is more likely to be correct than the previous determination. Certainly better-sampled light curves of longer duration would yield a more definitive result." }, "0808/0808.2391_arXiv.txt": { "abstract": "We present an analysis of the radio properties of large samples of Lyman Break Galaxies (LBGs) at $z \\sim 3$, 4, and 5 from the COSMOS field. The median stacking analysis yields a statistical detection of the $z \\sim 3$ LBGs (U-band drop-outs), with a 1.4 GHz flux density of $0.90 \\pm 0.21 \\mu$Jy. The stacked emission is unresolved, with a size $< 1\"$, or a physical size $< 8$kpc. The total star formation rate implied by this radio luminosity is $31\\pm 7$ $M_\\odot$ year$^{-1}$, based on the radio-FIR correlation in low redshift star forming galaxies. The star formation rate derived from a similar analysis of the UV luminosities is 17 $M_\\odot$ year$^{-1}$, without any correction for UV dust attenuation. The simplest conclusion is that the dust attenuation factor is 1.8 at UV wavelengths. However, this factor is considerably smaller than the standard attenuation factor $\\sim 5$, normally assumed for LBGs. We discuss potential reasons for this discrepancy, including the possibility that the dust attenuation factor at $z \\ge 3$ is smaller than at lower redshifts. Conversely, the radio luminosity for a given star formation rate may be systematically lower at very high redshift. Two possible causes for a suppressed radio luminosity are: (i) increased inverse Compton cooling of the relativistic electron population due to scattering off the increasing CMB at high redshift, or (ii) cosmic ray diffusion from systematically smaller galaxies. The radio detections of individual sources are consistent with a radio-loud AGN fraction of 0.3\\%. One source is identified as a very dusty, extreme starburst galaxy (a 'submm galaxy'). ", "introduction": "The power of discovering high redshift galaxies via the broad-band drop-out technique, ie. Lyman Break galaxies (LBGs), is now well established (Steidel et al. 1999; 2000). Using this technique, over a thousand galaxies have now been detected at $z > 2$. Detailed studies show that these galaxies have stellar masses between 10$^{10}$ and 10$^{11}$ $M_\\odot$, and a (comoving) volume density at $z \\sim 3$ of $\\sim 0.005$ Mpc$^{-3}$ (Giavalisco 2002). A number of issues are still being investigated for LBGs. One important issue is deriving the total star formation rate. Besides model parameters such as the IMF and star formation history, the dust correction in the UV remains under investigation. Steidel et al. (1999) originally estimated a typical UV dust attenuation factor of about 5, based on optical spectroscopy. Adelberger \\& Steidel (2000) applied a similar method to a larger sample of LBGs, as well as observations at other wavebands (radio, submm), and conclude: \"...the mean extinction at 1600 \\AA ~ for LBGs, a factor of 6 in our best estimate, could lie between a factor 5 and a factor of 9.\" More recently, Reddy \\& Steidel (2004) have derived the average dust correction factors for UV-selected galaxies at $z \\sim 2$ based on deep X-ray, radio, and optical spectroscopic studies of galaxies in the GOODS north field. They find that, for galaxies with total star formation rates $> 20$ $M_\\odot$ year$^{-1}$, the dust attenuation factor is between 4.4 and 5.1. A second issue for LBGs is the AGN fraction. Shapley et al. (2003) found that $\\sim 3\\%$ of LBGs at z$\\sim$3 have optical emission line spectra that are consistent with AGN. The Cosmic Evolution Survey (COSMOS), covering 2 \\sq\\deg, is a comprehensive study of the evolution of galaxies, AGN, and dark matter as a function of their cosmic environment. The COSMOS field has state-of-the-art multiwavelength observations, ranging from the radio through the X-ray (Scoville et al. 2007; Capak et al. 2007). A major part of this study is to identify the largest samples of LBGs to date. In section 2 we describe the LBG samples from the COSMOS survey based on U, B, and V-band drop-out searches. In this paper we consider the radio properties of these large samples of LBGs, similar to the study of $z = 5.7$ Lyman-$\\alpha$ emitting galaxies by Carilli et al. (2007). Observations of the COSMOS field have been done at 1.5$\"$ resolution (FWHM) at 1.4 GHz with the Very Large Array (Schinnerer et al. 2007). We employ the new VLA image that has added integration time in the central $\\sim 1^o$ (Schinnerer in prep.), giving an rms at the field center of $\\sim 7$ $\\mu$Jy beam$^{-1}$. In section 3.1 we search for radio counterparts to individual LBGs, while in section 3.2 we perform a stacking analysis to derive the statistical properties of the samples. In section 4 we discuss the implications these observations have on the questions of dust extinction and the AGN fraction in LBGs. We adopt a standard concordance cosmology. ", "conclusions": "\\subsection{The AGN fraction and massive starbursts} At the sensitivity limits of current deep radio surveys, such as COSMOS, detection of individual sources at $z > 2$ is limited to either extreme starbursts (star formation rates $> 1000$ $M_\\odot$ year$^{-1}$), or radio jet sources with luminosities comparable to M87. After correcting for random detections, we obtain a detection rate of $\\sim 0.3\\%$. Shapley et al. (2003) find an AGN fraction in LBG samples of $\\sim 3\\%$, based on optical spectroscopy. If all of our detected sources are radio AGN, then the implied radio loud fraction is $\\sim 10\\%$. Interestingly, a value of 10\\% is the canonical value for radio loud AGN based on studies of nearby galaxies (eg. Ivezic et al. 2002; Petric et al. 2007). We emphasize caution in interpreting our results on the radio loud fraction of AGN at high redshift, for a number reasons. First, in a recent comparison of the FIRST and SDSS surveys, Jiang et al. (2007) find that the radio loud fraction likely depends on both optical luminosity and redshift. This study is not directly comparable to our COSMOS study, since they were limited to much more luminous sources at high redshift, by about two orders of magnitude in the radio. Second, the above analysis does not consider the possibility that some of the 0.3\\% of the radio detections are extreme, highly obscured starbursts, comparable to the submm galaxies. Indeed, in a companion paper, we present the discovery of a submm galaxy in our radio-selected LBG COSMOS sample (Capak et al. 2008). The radio-detected LBG has a spectroscopic redshift of $z =4.5$, and a radio flux density of $45\\pm 10\\mu$Jy. Thermal emission from warm dust is detected from this galaxy at 250GHz using the MAMBO bolometer camera, with a flux density of $3.4\\pm 0.7$mJy. This source is the first submm galaxy yet identified (spectroscopically) at $z > 4$ (Capak et al. 2008). And third are the caveats mentioned in section 2 concerning our conservative source selection criteria, eg. excluding obvious optically bright broad line AGN. \\subsection{The sub-$\\mu$Jy radio source population} The unprecedented size of the COSMOS LBG sample has allowed us to reach sub-$\\mu$Jy sensitivity levels in the stacking analysis of the LBG samples. We obtain a clear statistical detection of $S_{\\rm 1.4GHz} = 0.90\\pm 0.21\\mu$Jy for the U-band ($z \\sim 3$) drop-outs. The implied rest frame 1.4 GHz luminosity density is $L_{\\rm 1.4\\rm GHz} = 5.1\\times 10^{29}$ erg s$^{-1}$ Hz$^{-1}$, assuming a spectral index of $-0.8$ (Condon 1992). Assuming the radio emission is driven by star formation, we can derive a total star formation rate (0.1 to 100 $M_\\odot$), from the rest frame 1.4 GHz luminosity density. There are a number of recent calibrations of this relationship for nearby galaxies using, eg. the IRAS and NVSS surveys. These studies adopt the star formation rates derived from the far-IR emission based on the relations in Kennicutt (1998), assuming a Salpeter IMF, and then calibrate the radio conversion factor using the tight radio to far-IR correlation for star forming galaxies. We adopt the conversion factor from the study of Yun, Reddy, and Condon (2001): $$\\rm SFR = 5.9 \\pm 1.8 \\times 10^{-29} {\\sl L}_{\\rm 1.4GHz} ~ M_\\odot~ year^{-1}, $$ \\noindent where $L_{\\rm 1.4\\rm GHz}$ is in erg s$^{-1}$ Hz$^{-1}$. From this, we calculate a mean star formation rate of $31\\pm 7$ $M_\\odot$ year$^{-1}$. Note that the conversion factor above is within 10\\% of that derived by Bell (2003). The mean observed UV luminosity at a rest frame wavelength of 1600 \\AA ~of the U-band drop-out sample is: $L_{2000A} = 1.2\\times 10^{29}$ erg s$^{-1}$ Hz$^{-1}$ (Capak et al. 2008 in prep). Using equation (1) in Kennicutt (1998) for the relationship between UV luminosity and total star formation rate, we derive a total star formation rate of 17 $M_\\odot$ year$^{-1}$, uncorrected for dust attenuation. The ratio of radio derived star formation rate to UV derived star formation rate is $1.8\\pm 0.4$. The simplest conclusion is that the UV emission is attenuated by dust by a factor of 1.8. However, this factor is considerably smaller than the standard factor $\\sim 5$ adopted for LBGs (see Section 1). We consider some possible reasons for this difference. One possibility is that the dust attenuation factor for the $z \\sim 3$ COSMOS LBG sample is indeed smaller than for other LBG samples. Most of the studies of dust attenuation of LBG galaxies have been at $z < 3$ (see section 1), and hence it is possible that at higher redshift the dust attenuation factor decreases. An argument against this decrease is the recent X-ray and optical study of a sample of LBGs at $z \\sim 3$ by Nandra et al. (2002), who also derive a dust attenuation factor of about 5. We also re-emphasize that the COSMOS LBG sample analyzed herein de-selected sources with poor phot-z fits, which can occur for very heavily obscured objects (Section 2). On the other hand, Wilkins, Trentham, \\& Hopkins (2008) present a detailed comparison of the build up of stellar mass in galaxies versus the cosmic star formation rate density. They conclude that there is a discrepancy between the rate of stellar mass creation, and the star formation rates derived from the UV luminosities, at high redshift, if one assumes the standard factor $\\sim 5$ dust attenuation. The discrepancy is in the sense that the UV derived star formation rates are too high. They find that the peak in the discrepancy occurs at $z \\sim 3$, and they suggest that the UV-derived star formation rates at this redshift may be over-estimated by a factor $\\sim 4$. They also point out that '...this large deviation at high redshift offers an explanation for why the integrated star formation history implies a local stellar mass density in excess of that measured.' A second possibility is that the conversion factor of radio luminosity to star formation rate is higher in the $z \\sim 3$ LBG COSMOS sample than has been derived for low redshift galaxies. Radio studies of 24$\\mu$m selected galaxies by Appleton et al. (2004) imply that the radio conversion factor is constant out to $z \\sim 1$, ie. that the radio-FIR correlation is constant out to this redshift. Reddy \\& Steidel (2004) extend this conclusion out to $z \\sim 2$ in their extensive study of UV selected galaxies. Most recently, Ibar et al. (2008) conclude that the radio conversion factor remains constant out to $z \\sim 3$, using a 24$\\mu$m source sample with redshifts from the SXDF. However, given the need for individual source detections in the radio, their $z \\sim 3$ radio sources have star formation rates about two orders of magnitude larger than the stacking results presented herein, and hence a direct comparison is problematic. One physical reason why we might expect the radio conversion factor to diverge at the highest redshifts is increased relativistic electron cooling due to inverse Compton scattering off the cosmic microwave background (CMB). The ratio of relativistic electron energy losses due to synchrotron radiation, to energy losses due to inverse Compton radiation, equals the ratio of the energy density in the magnetic field to that in the photon field. The energy density in the CMB increases as: $U_{CMB} = 4.2\\times 10^{-13} (1+z)^4$ ergs cm$^{-3}$. Figure 2 shows a comparison of $U_{CMB}$ with the typical energy densities in the magnetic fields in different regions in galaxies. We show the range of fields considered typical for spiral arms ($\\sim$ few $\\mu$G), and for starburst galaxy nuclei (of order 100$\\mu$G; Beck et al. 1994; see review by Beck 2005). The important point is that IC losses off the CMB will dominate synchrotron losses in a typical ISM at $z \\ge 0.5$, and dominate in starburst nuclei at $z \\ge 4$. We note that inverse Compton losses will not affect the thermal electrons responsible for Free-Free emission from star forming galaxies. Such Free-Free emission may dominate the total radio emission from galaxies at rest frequencies between roughly 40 GHz and 100 GHz (Condon 1992), and hence become an important factor in radio continuum studies of very high redshift galaxies. A depressed radio luminosity for a given star formation rate could also arise if the $z \\sim 3$ LBGs are systematically smaller galaxies. The standard model that produces the radio-FIR correlation (Condon 1992) requires a cosmic ray processing box size $\\ge 1$ kpc. It has been observed that dwarf galaxies at low redshift depart from the radio-FIR correlation by about a factor of 2, in the sense of being radio under-luminous. The hypothesis is that the cosmic rays diffuse out of the galaxy on timescales shorter than required to maintain the standard radio-FIR correlation (Yun et al. 2001). Giavalisco (2002) describes high $z$ LBGs as having typical half-light radii of 4 to 7 kpc, significantly larger than typical dwarf galaxies. However, he points out that: \"..frequently the galaxies have disturbed or fragmented morphologies, with one bright core, or multiple knots embedded in diffuse nebulosity, reminiscent of merger events.\" In this paper, we report the first robust statistical (median) detection of sub-$\\mu$Jy radio emission from LBG galaxies at $z \\sim 3$. This detection was made possible by the very wide area, and depth, of the Cosmos field. While our physical interpretation of the result remains inconclusive, there are a number of future studies we are pursuing to address the interesting questions raised concerning the UV attenuation factor for LBGs, and the the radio luminosity to star formation rate conversion factor, at $z \\ge 3$. The most important study involves obtaining optical spectra of a large sample of LBGs from the COSMOS sample. Spectra will elucidate the nature of the sources, and allow for a study of the dust correction factor as a function of galaxy type (AGN, starburst, elliptical...). Also, the selection criteria for this LBG sample were very conservative (Section 2). We will further refine (and increase) our LBG samples using the Cosmos multiband photometry, and perform statistical analyses of the radio properties as a function of eg. stellar mass, or $A_V$. We are also exploring the IR properties of these samples with Spitzer. Such IR studies have particular relevance in the light of the results of Ivison et al. (2007), who find that for the seven IR luminous $z \\sim 3$ LBGs in their AEGIS sample ($S_{24\\mu m} > 60\\mu$Jy), the median 1.4 GHz flux density is 44$\\mu$Jy, comparable to that expected for sub-mm galaxies. Lastly, unambiguous identification of radio AGN using eg. optical spectra, will provide input into faint radio source population models that are being generated in the context of planning for future, large area radio telescopes (Wilman et al. 2008). The results presented herein indicate that future deep (sub-$\\mu$Jy), wide field surveys with the Expanded Very Large Array, will detect routinely the radio emission from individual LBGs out to $z \\sim 3$." }, "0808/0808.1939_arXiv.txt": { "abstract": "I decompose the unstable growing modes of stellar disks to their Fourier components and present the physical mechanism of instabilities in the context of resonances. When the equilibrium distribution function is a non-uniform function of the orbital angular momentum, the capture of stars into the corotation resonance imbalances the disk angular momentum and triggers growing bar and spiral modes. The stellar disk can then recover its angular momentum balance through the response of non-resonant stars. I carry out a complete analysis of orbital structure corresponding to each Fourier component in the radial angle, and present a mathematical condition for the occurrence of van Kampen modes, which constitute a continuous family. I discuss on the discreteness and allowable pattern speeds of unstable modes and argue that the mode growth is saturated due to the resonance overlapping mechanism. An individually growing mode can also be suppressed if the corotation and inner Lindblad resonances coexist and compete to capture a group of stars. Based on this mechanism, I show that self-consistent scale-free disks with a sufficient distribution of non-circular orbits should be stable under perturbations of angular wavenumber $m>1$. I also derive a criterion for the stability of stellar disks against non-axisymmetric excitations. ", "introduction": "\\label{sec:intro} Both $N$-body simulations \\citep{H71} and analytical methods (Kalnajs 1978; Jalali \\& Hunter 2005, hereafter JH) show that global instabilities can generate barred structures in galactic disks. Apart from the bar mode, which is an isolated event in frequency space, global spiral modes seen in the eigenspectra of cored stellar disks (Jalali 2007, hereafter Paper I) constitute a discrete family that bifurcates form stationary \\citet{vK55} modes. But not all spiral structures are global modes as disturbances induced by close neighbors \\citep{BH92} and density inhomogeneities \\citep{T90} may also create the spiral patterns of the observed galaxies. A mode of a stellar disk is a mathematical entity that comes out of an eigenvalue problem. However, its physical origin in isolated systems has not yet been understood clearly. Lynden-Bell \\& Kalnajs (1972, hereafter LBK) attempted to explain a mode through the transport of angular momentum between different parts of the disk. They suggested that the inner Lindblad resonance (ILR) releases the angular momentum of central regions and the spiral structure transports it to the outer parts through the corotation (CR) and outer Lindblad resonances (OLR). This mechanism is favored by some galactic dynamicists \\citep{ATHA03}, but it is seriously challenged by Toomre's (1981) theory that says that feedback through the galactic center is a critical ingredient for growing modes. In Toomre's theory, on the other hand, the modeling of feedback as the reflection of a leading spiral wave at the galactic center and its emergence as a trailing one, is a simple description of a very complicated dynamics that governs the motions of stars. Unresolved issues concerning the evolution of unstable modes include the following: (i) The swing amplification theory is not capable of predicting the fate of a growing mode against other stationary and unstable modes that coexist in the eigenspectrum of a given model. (ii) How does the bar mode saturate? \\citep{KJKJ07}. (iii) Why does the nonlinear bar terminate almost at the corotation radius? \\citep{S81} (iv) We should also understand the origin of different species in an eigenspectrum and interpret their continuous or discrete nature, and the distribution of their pattern speeds and growth rates. Stellar orbits in a galactic disk begin to evolve once the surface density deviates from its equilibrium state, $\\Sigma_0(\\textbf{\\textit{x}})$, and develops a time-varying mean-field potential $V_1(\\textbf{\\textit{x}},t)$. Here $\\textbf{\\textit{x}}$ denotes the position vector of stars at the time $t$. In the linear regime, we are usually interested in density waves that grow/decay according to the exponential law $e^{st}$ and rotate with the fixed pattern speed $\\Omega_p$. For a wave of $m$-fold symmetry, the perturbed potential becomes \\begin{equation} V_1=\\epsilon e^{st} \\tilde V \\left (\\textbf{\\textit{x}},m\\Omega_p t \\right ). \\label{eq:perturbed-potnetial-introduction} \\end{equation} Since we are dealing with infinitesimal perturbations, I have introduced the small parameter $\\epsilon$ so that $\\epsilon e^{st}\\ll 1$. From (\\ref{eq:perturbed-potnetial-introduction}) one arrives at the equations of motion \\begin{equation} \\dot \\textbf{\\textit{x}} = \\textbf{\\textit{v}},~~ \\dot \\textbf{\\textit{v}} = -\\frac{\\partial V_0}{\\partial \\textbf{\\textit{x}} } - \\epsilon e^{st} \\frac{\\partial \\tilde V}{\\partial \\textbf{\\textit{x}} }, \\label{eq:equation-motion-for-x-and-v} \\end{equation} where $V_0(\\textbf{\\textit{x}})$ is the equilibrium potential field generated by galactic stars and a possible dark matter halo. In writing equation (\\ref{eq:equation-motion-for-x-and-v}), I have assumed that the motion of stars is restricted to the disk plane. When the equilibrium state is axisymmetric and the dark component is spherical, $V_0$ becomes a function of radial distance to the galactic center and the unperturbed equations (with $\\epsilon=0$) are integrable. In such a circumstance, the phase space is filled by rosette orbits denoted by $[\\textbf{\\textit{x}}_0(t),\\textbf{\\textit{v}}_0(t)]$. The growth of perturbations, whatever the magnitude of $\\epsilon e^{st}\\ll 1$ may be, deforms stellar orbits. Orbital deformations are measured by $\\tilde \\textbf{\\textit{x}}=\\textbf{\\textit{x}}-\\textbf{\\textit{x}}_0$ and $\\tilde \\textbf{\\textit{v}}=\\textbf{\\textit{v}}-\\textbf{\\textit{v}}_0$, which can be used in (\\ref{eq:equation-motion-for-x-and-v}) to obtain \\begin{equation} \\frac{d\\tilde \\textbf{\\textit{x}} }{dt}=\\tilde \\textbf{\\textit{v}},~~ \\frac{d\\tilde \\textbf{\\textit{v}} }{dt}=- \\left [ \\frac{\\partial^2 V_0}{\\partial \\textbf{\\textit{x}}^2 } \\right ]_{ \\textbf{\\textit{x}}_0 } \\!\\!\\!\\! \\cdot \\tilde \\textbf{\\textit{x}} - \\epsilon e^{st} \\left [ \\frac{\\partial \\tilde V}{\\partial \\textbf{\\textit{x}} } \\right ]_{ \\textbf{\\textit{x}}_0 }. \\label{eq:equation-motion-for-tilde-x-and-v} \\end{equation} Although a proper equilibrium distribution function $f_0(\\textbf{\\textit{x}},\\textbf{\\textit{v}})$ can self-consistently reproduce $\\Sigma_0(\\textbf{\\textit{x}})$ using rosette orbits, the perturbed density $\\Sigma_1(\\textbf{\\textit{x}},t)$ (corresponding to $V_1$) cannot be supported by rosette orbits alone and orbital deformations are necessary for the self-consistency of density waves. According to equations (\\ref{eq:equation-motion-for-x-and-v}) and (\\ref{eq:equation-motion-for-tilde-x-and-v}), orbital deformations of ${\\cal O}\\left ( \\epsilon e^{st} \\right )$ are sufficient to support the growth of density/potential perturbations up to the same order of magnitude of such deformations over a time scale of $1/{\\cal O}\\left ( \\epsilon e^{st} \\right )$. As the time is elapsed, the amplitude of perturbations increases exponentially and the solution of the linearized collisionless Boltzmann equation (CBE) fails when $\\epsilon e^{st}\\sim 1$. During my mode calculations, I realized that the orbital axes of certain stars librate in a coordinate frame that rotates with the density pattern. This {\\it resonant capture} initially seemed to be a higher-order nonlinear effect but further experiments showed that the resonant gap is constrained by the magnitude of density perturbations. The complex behavior of stars for infinitesimally small yet non-zero $\\epsilon e^{st}\\ll 1$, and the role of resonant stars in the generation of discrete galactic modes, are investigated in this paper. I use the results of Paper I and introduce a new dynamical mechanism that sparks unstable modes and governs the singular oscillations of \\citet{vK55} modes. Resolving the origin of instabilities and amplitude saturation precede my nonlinear calculations, which were made feasible in Paper I by the Petrov-Galerkin method and reducing the CBE to a system of nonlinear ordinary differential equations. Those reduced equations, however, are valid only when orbits are regular and averaging over angle variables is allowed. As unstable modes grow, chaotic orbits come into existence and the weighted residual form of the CBE must be modified to handle them. I quote some of the results of such modifications in this paper when I discuss the issue of mode saturation. In Paper III, I will give a full account of the mathematical and numerical modeling of stochastic layers, and will analyze modal interactions after their saturation phase. For the cored exponential disk embedded in the field of the cored logarithmic potential, I describe the decomposed Fourier components of unstable bar and spiral modes in \\S\\ref{sec:mode-decomposition} and highlight the existence of a phase shift between different components. In \\S\\ref{sec:capture-into-resonances}, I derive a condition for the corotation of the orbital axes of an ensemble of stars and explain the role of such a synchronous motion in pattern formation. I dedicate \\S\\ref{sec:perturbed-stellar-dynamics} to exploring the orbital structure of a perturbed stellar disk and identify a resonance mechanism that can generate both stationary and growing modes. I reveal the mechanism of angular momentum transfer between Fourier components and derive analytical expressions for the growth of resonance zone. I address the origin of instabilities in \\S\\ref{sec:origin-of-instabilities}, present a saturation mechanism for unstable modes in \\S\\ref{sec:mode-saturation}, and discuss about the global stability of soft-centered and scale-free disks. I explain the restrictions of LBK's mode mechanism in \\S\\ref{sec:discussion-and-conclusions} and end up the paper with concluding remarks. ", "conclusions": "\\label{sec:discussion-and-conclusions} I had already given an evidence in \\S\\ref{sec:modes-in-frequency-space} that LBK's theory cannot be extended to large growth rates. Here I provide a more detailed analysis of LBK's mode mechanism. Taking the $s\\rightarrow 0$ limit of (\\ref{eq:dL-dt-from-JH}) leads to equation (30) of LBK. The resulting integrand of the $l$th component involves the term $\\delta(l\\Omega_R+m\\Omega_{\\phi}-\\omega)$ but its argument never becomes zero for $l\\le -1$ as long as $\\Omega_p$ exceeds $\\Omega_{\\rm ILR}(m)$. Once the delta function vanishes for $\\Omega_p>\\Omega_{\\rm ILR}(m)$, the angular momentum content of $l\\le -1$ components remains zero. Therefore, angular momentum transfer between inner and outer resonances as suggested by LBK, is feasible only for $\\Omega_p < \\Omega_{\\rm ILR}(m)$. This is a constraint on the applicability of LBK's theory. According to the arguments of \\S\\ref{sec:role-of-ILR} and the WKB theory, the ILR is not transparent to short wavelength disturbances with $X\\gg 1$ \\citep{BT08,Mark74} and any developing spiral structure must be damped unless the ILR remains invisible to the CR. Therefore, an inconsistent point in LBK's paper is the imagination of an angular-momentum-transferring spiral structure while the condition $\\Omega_p <\\Omega_{\\rm ILR}(m)$ has already abandoned the existence of such structures. An issue yet needs to be explained: If there is no spiral structure in stable disks, which mechanism does transfer the angular momentum between inner and outer resonances? In fact, in the limit of $s\\rightarrow 0$ the density components $\\Sigma^{l-}_1$ and $\\Sigma^{l+}_1$ become stationary waves that are azimuthally separated by a phase shift of $90^{\\circ}$ (like the components of mode B1 in Figure \\ref{pic:components-modes-B1-S2}). Thus, they exert opposite gravitational torques on each other without any need for a communicating spiral structure. In this paper I decomposed unstable modes of a model disk galaxy to its constituent Fourier components and showed how different components experience a gravitational torque. My results clearly showed that only the Fourier component associated with the CR generates a resonant zone in the phase space, and other components exchange angular momentum far from resonances. This result led me to introduce a new dynamical mechanism that triggers unstable modes. According to my calculations, an irreversible resonant capture of stars into the CR causes a synchronous precession of their orbital axes, which in turn, support a rotating density pattern. The emerged pattern grows because the resonant zone expands in the frequency space. The irreversibility of resonant trapping is the most destructive event that happens in a stellar disk when a group of resonant stars with $d_i1$ the emergence of an ILR is inevitable in scale-free models. The competition between the CR and ILR can therefore stabilize sufficiently warm scale-free disks against $m>1$ excitations. \\citet{ER98b} suggested in \\S6.1 of their paper that self-consistent scale-free disks (without inner cutouts) do not admit growing non-axisymmetric modes at all, and they ruled out the possibility of a critical temperature. Their prediction is in agreement with my results except in two aspects: (i) Since the line $\\mu^{-1}_{1}({\\bf \\Omega})=0$ does not intersect the frequency space of scale-free models, there is no ILR to compete with a growing CR and density waves will be amplified for $m=1$. (ii) The ILR in cold disks is very special and it cannot develop a resonant zone of finite size. It is therefore hard to believe a serious influence by the ILR on the phase space flows up to a critical temperature. Note that each unstable mode dominates an isolated region of the frequency (action) space until a chaotic layer emerges due to resonance overlapping. The rate by which the chaotic layer diffuses itself in the phase space, is determined by the resistance of KAM tori against competing resonant zones. The limited space that the fastest growing bar mode occupies after its saturation (Khoperskov et al. 2007), is an $N$-body evidence for such a resistance of spiral modes that fill outer regions of the cored exponential disk. The other implication of resonance overlapping, which has observational support too, is that the central bar in grand-design barred spiral galaxies must join the spiral arms at the tips of the bar, which are the overlapping regions of two neighboring B and S modes in the frequency space (see Figure \\ref{pic:zones-overlap}). I note that the pattern speeds of the spiral and bar components are not the same because these structures are associated with different modes. The overlapping region is filled by chaotic orbits that yield the turbulent mixing of density waves. This process can enhance star formation near the tips of the bar. If we accept this scenario, barred spiral galaxies should have been formed from hot stellar disks whose eigenfrequency spectra include few spiral modes. On the other hand, flocculent spirals with many wave packets will be the natural destiny of initially cold disks that give birth to a rich family of spiral modes. The mathematical background for the nonlinear evolution of modes was developed in Paper I but that formulation is valid as long as averaging is allowed over angle variables. In the presence of chaotic orbits it is impossible to average out resonant angles and some modifications are required to deal with the CBE in its full nonlinear form. I will present such modifications in Paper III with the aim of discovering the dynamical processes that generate different classes of barred and spiral structures." }, "0808/0808.3611_arXiv.txt": { "abstract": "Using the 100-m radio telescope at Effelsberg, we mapped a large area around the Andromeda Galaxy in the 21-cm line emission of neutral hydrogen to search for high-velocity clouds (HVCs) out to large projected distances in excess of $100~\\mathrm{kpc}$. Our $3 \\, \\sigma$ \\ion{H}{i} mass sensitivity for the warm neutral medium is $8 \\times 10^{4}~{\\rm M}_{\\odot}$. We can confirm the existence of a population of HVCs with typical \\ion{H}{i} masses of a few times $10^{5}~{\\rm M}_{\\odot}$ near the disc of M31. However, we did not detect any HVCs beyond a projected distance of about $50~\\mathrm{kpc}$ from M31, suggesting that HVCs are generally found in proximity of large spiral galaxies at typical distances of a few $10~\\mathrm{kpc}$. Comparison with CDM-based models and simulations suggests that only a few of the detected HVCs could be associated with primordial dark-matter satellites, whereas others are most likely the result of tidal stripping. The lack of clouds beyond a projected distance of $50~\\mathrm{kpc}$ from M31 is also in conflict with the predictions of recent CDM structure formation simulations. A possible solution to this problem could be ionisation of the HVCs as a result of decreasing pressure of the ambient coronal gas at larger distances from M31. A consequence of this scenario would be the presence of hundreds of mainly ionised or pure dark-matter satellites around large spiral galaxies like the Milky Way and M31. ", "introduction": "One of the currently most favoured cosmological models is the so-called Lambda Cold Dark Matter ($\\Lambda$CDM) model. It assumes that the evolution of the universe is dominated by dark energy and dark matter which in sum account for 96~per~cent of the total energy density of the universe \\citep{Spergel2003}. An important prediction of CDM models is the hierarchical formation of gravitationally bound structures. The smallest dark-matter haloes are expected to form first, whereas larger structures, ranging from spiral galaxies to galaxy clusters, are formed at a later stage through merging and accretion of smaller dark-matter haloes (bottom-up scenario). Numerical simulations of structure formation in CDM cosmologies have successfully reproduced the mass function and radial distribution of galaxies on the scales of galaxy clusters. On smaller scales, however, simulations predict significantly more dark-matter haloes than being observed \\citep{Klypin1999,Moore1999}. This discrepancy has been named the `missing satellites' problem. To overcome this problem, \\citet{Blitz1999} suggested that high-velocity clouds (HVCs) might be the gaseous counterparts of the `missing' dark-matter haloes around the Milky Way. HVCs are gas clouds observed all over the sky in the 21-cm line emission of neutral atomic hydrogen. They were discovered with the Dwingeloo radio telescope by \\citet{Muller1963}, and they are characterised by high radial velocities of typically $|v_{\\rm LSR}| \\gtrsim 100~\\mathrm{km \\, s}^{-1}$ (see \\citealt{Wakker1991a} for details). If HVCs were the `missing' satellites, they could not have experienced significant star formation during their evolution. Therefore, they would appear in the form of pure gas clouds without any noticeable stellar population. In addition, HVCs would be spread all over the Local Group with typical distances of hundreds of kpc and high \\ion{H}{i} masses of about $10^7~{\\rm M}_{\\odot}$. At the same time, the expected large distances of HVCs from the Milky Way would result in fairly decent angular diameters of the clouds. Therefore, \\citet{Braun1999} defined a sub-sample of compact and isolated HVCs (the so-called CHVCs) which are characterised by angular sizes of less than $2^{\\circ}$~FWHM as well as isolation and separation from neighbouring \\ion{H}{i} emission. The overall kinematics of the CHVC population is consistent with a distribution throughout the Local Group \\citep{deHeij2002b}, making them promising candidates for the `missing' dark-matter satellites around the Milky Way and the Andromeda Galaxy. \\citet{Putman2002} extended the CHVC catalogue into the southern hemisphere, using the \\ion{H}{i} Parkes All-Sky Survey \\citep{Barnes2001}. The data for the northern and southern hemispheres were later combined by \\citet{deHeij2002b} into an all-sky catalogue of 216~CHVCs. Several arguments have been raised against the idea of HVCs and CHVCs being the `missing' dark-matter haloes predicted by CDM cosmologies. First of all, attempts to identify a population of HVCs in other nearby galaxy groups \\citep{Zwaan2001,Braun2001,Pisano2004} have failed, resulting in upper distance limits for HVCs and CHVCs from the Milky Way of the order of $150~\\mathrm{kpc}$. Additional evidence for HVCs being nearby at distances of the order of only $10~\\mathrm{kpc}$ has been provided by the detection of H$\\alpha$ emission from both HVCs and CHVCs (e.g., \\citealt{Kutyrev1986,Weiner2001,Tufte2002,Putman2003}) and from the determination of distance brackets for several HVC complexes (e.g., \\citealt{Danly1993,Wakker1996,Wakker2007a,Wakker2007b,Thom2006,Thom2008}). Small distances from the Milky Way are also consistent with the head-tail structures found in numerous HVCs and CHVCs (e.g., \\citealt{Bruens2000,Bruens2001,Westmeier2005b}) which are thought to result from ram-pressure interaction of the clouds with the ambient gas of the Galactic corona \\citep{Quilis2001,Konz2002}. \\begin{table} \\caption{Observational parameters of the northern and southern part of our Effelsberg \\ion{H}{i} blind survey of M31. The final baseline RMS is given for a system temperature of $25~\\mathrm{K}$ at the given velocity resolution. The specified sensitivity is the baseline RMS times the spectral bin width at the original velocity resolution, converted to \\ion{H}{i} column density. Sensitivities for the WNM were calculated for a brightness temperature of $T_{\\rm B} > 3 \\, \\sigma$, a velocity resolution of $20 \\; \\mathrm{km \\, s}^{-1}$, and a line width of $25 \\; \\mathrm{km \\, s}^{-1}$~FWHM.} \\label{tab_obspar} \\begin{center} \\begin{tabular}{lrrl} \\hline parameter & south & north & unit \\\\ \\hline autocorrelator & old & \t new & \t\t\t \\\\ bandwidth & 6.3 & \t 10 & MHz\t\t\t \\\\ polarisations & \t 2 & \t 2 & \t\t\t \\\\ channels per polarisation & 512 & \t4096 & \t\t\t \\\\ velocity resolution & 2.6 & \t 0.5 & $\\mathrm{km \\, s}^{-1}$ \\\\ frequency switching & normal & in-band & \t\t\t \\\\ total integration time & 180 & \t 180 & s\t\t\t \\\\ final baseline RMS & \t45 & \t 60 & mK\t\t\t \\\\ sensitivity & \t21 & \t 5.5 & $10^{16} \\; \\mathrm{cm}^{-2}$ \\\\ WNM sensitivity & 2.2 & \t 1.3 & $10^{18} \\; \\mathrm{cm}^{-2}$ \\\\ WNM mass sensitivity & \t 8 & \t 5 & $10^4 \\; {\\rm M}_{\\odot}$\t \\\\ \\hline \\end{tabular} \\end{center} \\end{table} The problem of determining the spatial distribution of HVCs can ultimately be solved by searching for the expected HVC population around the nearest large spiral galaxy, the Andromeda Galaxy (M31). First, the distance of M31 is well known so that important physical parameters of HVCs, such as their \\ion{H}{i} mass or their diameter, can directly be determined from the observations. Second, we will look at the HVC population of M31 from the outside, allowing us to determine the (projected) radial distribution of HVCs. Finally, M31 is relatively close to the Milky Way. Therefore, the sensitivity and angular resolution of large single-dish radio telescopes will easily allow us to detect HVCs of the mass and size of the large HVC complexes observed near the Milky Way. The first comprehensive search for HVCs around M31 was carried out by \\citet{Thilker2004} with the 100-m Green Bank Telescope. They mapped an area of $7^{\\circ} \\times 7^{\\circ}$ in the 21-cm line of \\ion{H}{i} with high sensitivity and discovered a population of about 20~HVCs out to the edge of their map at about $50~\\mathrm{kpc}$ projected distance from M31. Their discovery marked the first detection of an extensive HVC population around a galaxy other than our own. Several of these HVCs were studied in detail with the Westerbork Synthesis Radio Telescope (WSRT) by \\citet{Westmeier2005a}. Thus, the radial extent of the population of HVCs and CHVCs around galaxies like the Milky Way and M31 is confined by two limits. An upper limit of about $150~\\mathrm{kpc}$ can be derived from the non-detection of HVCs in other galaxy groups, and a lower limit of about $50~\\mathrm{kpc}$ is marked by the edge of the \\ion{H}{i} survey of \\citet{Thilker2004} out to which HVCs were found to be present. Therefore, we decided to carry out a complementary \\ion{H}{i} survey with the 100-m radio telescope at Effelsberg to search for HVCs and CHVCs around M31 out to much larger projected distances in excess of $100~\\mathrm{kpc}$. This would allow us to trace the distribution of the HVC population of M31 over its entire radial extent and compare our results with the predictions of CDM structure formation scenarios. Our paper is organized as follows. In Sect.~\\ref{sect_observations} we explain the technical aspects of our observations. In Sect.~\\ref{sect_reduction} our data reduction, calibration, and analysis strategy is described. In Sect.~\\ref{sect_completeness} we discuss the various completeness issues of our survey and their implications for our results. In Sect.~\\ref{sect_results} we describe the derived observational and physical properties of the high-velocity clouds detected in our survey. In Sect.~\\ref{sect_comparison} we discuss the results of our comparison of the observational parameters of the HVCs near M31 with different CDM-based models and with the distribution of satellite galaxies around M31. Sect.~\\ref{sect_origin} discusses the evidence for different hypotheses on the origin of HVCs. Finally, Sect.~\\ref{sect_summary} summarises our results and conclusions. \\begin{figure*} \\includegraphics[width=0.74\\linewidth]{westmeier2.eps} \\caption{The plot shows the geometry of the field (thick solid line) mapped around M31. The dashed circles separate the five different regions (labelled with numbers) for which individual spatial completeness calculations have to be performed.} \\label{fig_spatcomp} \\end{figure*} ", "conclusions": "\\label{sect_summary} We used the 100-m radio telescope at Effelsberg to map a large area around the Andromeda Galaxy, M31, in the 21-cm line emission of neutral atomic hydrogen. Our survey extends out to a projected distance of about $140~\\mathrm{kpc}$ in the south-eastern direction (equivalent to about two thirds of the projected distance towards M33) and about $70~\\mathrm{kpc}$ in the north-western direction. With this map layout we are able to fill the previously existing gap between the outer boundary of the GBT survey of \\citet{Thilker2004} at about $50~\\mathrm{kpc}$ projected distance from M31 and the upper limit of about $150~\\mathrm{kpc}$ for the distance of HVCs as derived from the non-detections in nearby galaxy groups \\citep{Zwaan2001,Braun2001,Pisano2004}. The achieved spectral baseline RMS is $45~\\mathrm{mK}$ at $2.6~\\mathrm{km \\, s}^{-2}$ velocity resolution, corresponding to a $3 \\, \\sigma$ \\ion{H}{i} column density detection limit of $2.2 \\times 10^{18}~\\mathrm{cm}^{-2}$ for the warm neutral medium ($\\Delta v = 25~\\mathrm{km \\, s}^{-1}$ FWHM). This translates into an \\ion{H}{i} mass sensitivity of $8 \\times 10^{4}~{\\rm M}_{\\odot}$. In total, we detected 17~individual HVCs and several regions of more extended extra-planar gas all around M31. The discrete clouds are predominantly unresolved by the HPBW of $9~\\mathrm{arcmin}$ and characterised by typical \\ion{H}{i} masses of a few times $10^{5}~{\\rm M}_{\\odot}$. We did not detect any clouds beyond a projected distance of about $50~\\mathrm{kpc}$, suggesting that HVCs are generally found in proximity of their host galaxies. In particular, we did not find an extended populations of hundreds of CHVCs as observed around the Milky Way, suggesting that the Galactic CHVCs are intrinsically small clouds in the immediate vicinity of the Milky Way. A comparison with the Local Group population model of CHVCs, as proposed by \\citet{deHeij2002b}, reveals that their best-fitting model {\\#}9 is discarded by our data with high confidence. Neither the observed projected radial distribution nor the \\ion{H}{i} mass function of HVCs around M31 can be explained by the model. Instead, we find that a Gaussian radial scale length of the order of $50~\\mathrm{kpc}$ can best explain the observed projected distribution of HVCs around M31. In addition, we find that the HVCs are also distinct from the M31 satellite galaxies through typically lower \\ion{H}{i} masses and smaller projected distances from M31. CDM-based structure formation simulations by \\citet{Kravtsov2004} suggest that about 50 to 100~dark-matter haloes with total gas masses of greater than $10^{6}~{\\rm M}_{\\odot}$ should exist within about $300~\\mathrm{kpc}$ of M31. Only~2 to~5 of these haloes should be located within $50~\\mathrm{kpc}$ of M31, suggesting that some of the HVCs found near M31 could instead be tidally stripped gas from present or former satellite galaxies of M31. This idea is supported by our high-resolution follow-up observations of several HVCs \\citep{Westmeier2005a}. The lack of detections beyond $50~\\mathrm{kpc}$ projected radius, however, is in conflict with the predictions made by \\citet{Kravtsov2004}. A possible explanation of this discrepancy could be ionisation as a result of decreasing pressure of the ambient coronal medium at larger distances from M31, as suggested by hydrostatic simulations of \\citet{Sternberg2002}. An important consequence of this scenario would be the presence of hundreds of mainly ionised or pure dark-matter satellites near large spiral galaxies, such as the Milky Way and M31, which would be undetectable in the 21-cm line of neutral hydrogen. Finding these almost invisible satellites would be an important observational result in support of CDM cosmologies. Another promising scenario recently discussed by \\citet{Bekki2008} is a possible tidal interaction between M31 and M33 during their previous encounter. This would have stripped gas from M33 which could have formed some of the HVCs observed near M31. More detailed and extensive observations and simulations will be required to further investigate this interesting scenario. With our Effelsberg \\ion{H}{i} survey of M31 we have shown that HVCs are most likely concentrated around the large spiral galaxies with typical distances of no more than a few $10~\\mathrm{kpc}$. It is likely that different physical processes, such as tidal stripping, accretion of primordial dark-matter haloes, or galactic outflows, have contributed to the HVC populations of M31 and the Milky Way. The general volatility of these processes could naturally explain some of the differences between the HVC populations of M31 and the Milky Way, for example the extended and complex gaseous streams of the Magellanic Clouds which have no counterpart in M31." }, "0808/0808.3561_arXiv.txt": { "abstract": "The stellar asymmetry of faint thick disk/inner halo stars in the first quadrant ({\\it l} = 20 --45$\\arcdeg$) first reported by \\cite{lar96} and investigated further by \\cite{par03,par04} has recently been confirmed by the SDSS \\citep{jur08}. Their interpretation of the excess in the star counts as a ringlike structure, however, is not supported by critical complementary data in the fourth quadrant, not covered by the SDSS. We present stellar density maps from the Minnesota Automated Plate Scanner (MAPS) Catalog of the POSS I showing that the overdensity does not extend into the fourth quadrant. The overdensity is most probably not a ring. It could be due to interaction with the disk bar, evidence for a triaxial thick disk, or a merger remnant/stream. We call this feature the Hercules Thick Disk Cloud. ", "introduction": "\\cite{lar96} initially reported a substantial asymmetry of faint blue stars in the first quadrant (Q1) of the inner Galaxy, $l = 20\\arcdeg - 45\\arcdeg$ compared with complementary fields in the fourth quadrant (Q4) based on star counts from MAPS\\footnote{The Minnesota Automated Plate Scanner Catalog of the POSS I is online at: http://aps.umn.edu} \\citep{cab03}. \\cite{par03} made a more in-depth survey to map the extent of the asymmetry using 40 contiguous fields in each of three regions: Q1 above and below the plane and Q4 above the plane. Q4 below the plane is not covered in the POSS I. They found a 25\\% excess in the number of probable thick disk stars in Q1 above \\it{and }\\rm below the plane when compared to the complementary Q4 fields. The region was irregular in shape and covered several hundred square degrees, but with a completeness limit at $\\approx$ 18--18.5 mag, the stars showing the excess in Q1 were relatively nearby, $\\sim$ 1 -- 2 kpc from the Sun. \\cite{par04} also found an associated kinematic signature, a significant lag of ~80 to 90 km/sec in the direction of Galactic rotation for the associated thick disk stars in Q1. The recent release of the SDSS Data Release 5 (DR5) photometry in the direction of the observed asymmetry in Q1 led to the discovery of a feature at much fainter magnitudes, the distant Hercules-Aquila cloud \\citep{bel07} and the photometric parallax study by \\cite{jur08} confirmed our nearer asymmetry in the inner Galaxy as an overdensity at a galactocentric radius of 6.5 kpc situated 1.5 kpc above the plane. The SDSS survey however is not well designed for a good study of the thick disk inside the Solar orbit. It extends below $b = 30\\arcdeg$ in only a few directions in Q1 and has only limited coverage in Q4. We are continuing our program of photometric and spectroscopic observations to map the size and extent of the asymmetry along our line of sight and to determine the degree of spatial and kinematic asymmetry above and below the plane. We are using the SMARTS Consortium CTIO 1-meter Y4KCam and the Steward Observatory 90\" Bok telescope 90Prime Mosaic imager to obtain wide-field multi-color CCD imaging to fainter completeness limits than the POSS I. Spectra of the candidate thick disk stars for radial velocities and metallicity estimates have been observed using the Hydra multi-object spectrometer on the CTIO Blanco 4-meter telescopes and the Hectospec on the MMTO 6.5 meter. In this Letter we present a stellar number density map we created from the MAPS POSS I scans to provide a more global reference for our current deeper but more spatially restricted photometric and spectroscopic study of the thick disk in the inner Galaxy. Other works \\citep{xu07} have used plate data to supplement gaps in the SDSS at more southern declinations with good success. Our map of the stellar density in Q1 and Q4 covers much of the sky unavailable to SDSS and demonstrates that the nearby asymmetry in Q1 does not represent a ring above the Galactic plane, but instead is a significant substructure or cloud extending over many square degrees in galactic longitude and latitude. We call this feature the Hercules Thick Disk cloud. ", "conclusions": "After the reduction steps outlined above, we then binned the stars $0.25 \\arcdeg \\times 0.25 \\arcdeg$ in {\\it l} and {\\it b} to create maps of the stellar density distribution of the faint blue and intermediate color stars shown in Figures~\\ref{fig5} and ~\\ref{fig6}. The figures are color-coded with respect to number density per 0.0625 square degree. \\cite{jur08} described the excess in Q1 as due to a ``ringlike\" structure because the overdensity appeared to be radially constant and circular in cross section in their Figure 27. The center of the overdensity region ranges from $(X,Y,Z)=(6.5 kpc, -2.2 kpc, 1.5 kpc)$ to $(6.5 kpc, 0.3 kpc, 1.5 kpc)$ and can easily be converted into galactic coordinates. This is shown as the purple line on Figure~\\ref{fig5}. If the feature were symmetric about $l=0\\arcdeg$ it would project into Q4 to $(6.5 kpc, 2.2 kpc, 1.5 kpc)$. This projection is also shown as a purple line on Figure~\\ref{fig6}. Comparison of Figures~\\ref{fig5} and ~\\ref{fig6}, however show that the density of these stars is not symmetric with respect to the Sun-Center line. There is a clear excess of stars in Q1 over Q4 in the range {\\it l} $= 25$ -- $45\\arcdeg$ and {\\it b} $= 30$ -- $40\\arcdeg$. Furthermore, we emphasize that the excess would not have been initially discovered if it had been a symmetric ring since we \\citep{lar96} were comparing star counts for complementary fields in Q1 and Q4. Could the ``ringlike\" structure be inclined to the plane and therefore not visible in Q4? Most probably not. The full height of the overdensity in Z over an azimuthal distance of 2500 parsecs is only 500 pc (Juric{\\c'} et al.). Even for the pathological case of a paper thin inclined distribution in Z the maximum inclination could only be 11 degrees and given its width in X it should have been visible in Figure~\\ref{fig6}. Additionally, there is strong evidence from Juri{\\'c} et al.'s Figure 27 (left panel) that for X = 7250-7750 parsecs the overdensity is exclusively above $Z = 1500$ pc. Examination of the same figure's middle panel shows that all significant contributors to the overdensity in this same X range have $Y < 1000$ pc. This would be the opposite of what should be happening if the overdensity were falling into the plane in Q4. Finally, \\cite{par04} studied the velocities of samples of stars taken from overdensity regions in Q1 compared with a control sample Q4 and found in a somewhat weak result that the Z component of velocity was less negative for Q1 than it was for Q4. If a coherent population of stars were moving together towards the disk from above the plane as they entered Q4, the opposite should be true. The cloud of stars detected by Juri{\\'c} et al. is not symmetric about the $l=0\\arcdeg$ line and almost certainly is not a ring. This does not change their other possible explanation for the feature, however. Given the broad extent of the cloud (Figures~\\ref{fig5} and ~\\ref{fig6}) together with its apparent small ranges in radial distance from the Sun and its distance above the galactic plane (their Figure 27), the Hercules Thick Disk cloud may be a debris stream consistent with the disk formation scenario described by \\cite{abd03}. While the Hercules Thick Disk Cloud is relatively nearby on the sky it is not related to the more distant (10 - 20 kpc) Hercules-Aquila cloud of \\citep{bel07}. The northern extent of the Hercules-Aquila cloud (Figure 2 in \\citet{bel07}) is confined to galactic latitudes less than $30\\arcdeg$ and even in those regions the bulk of the stars are much fainter than our magnitude limit. The contamination of our sample by Hercules-Aquila cloud stars would be less than 5 objects per square degree (0.5 stars/bin) in any case given our relatively bright magnitude limits. Other explanations are still possible such as a triaxial thick disk and an overdensity due to an interaction with the disk bar. Analysis of our CCD photometry and spectroscopy for fainter stars in Q1 and Q4 will be used to address this question." }, "0808/0808.0081_arXiv.txt": { "abstract": "In this work we estimate the radius and the mass of a self-gravitating system made of axions. The quantum axion field satisfies the Klein-Gordon equation in a curved space-time and the metric components of this space-time are solutions to the Einstein equations with a source term given by the vacuum expectation value of the energy-momentum operator constructed from the axion field. As a first step towards an axion star we consider the up to the $\\phi^6$ term in the axion potential expansion. We found that axion stars would have masses of the order of asteroids ($\\sim 10^{-10}M_\\odot$) and radius of the order $\\sim \\mbox{few} ~$centimeters. ", "introduction": " ", "conclusions": "" }, "0808/0808.1805_arXiv.txt": { "abstract": "The brick wall method in calculations of the entropy of black holes can be applied to the FRW cosmology in order to study the statistical entropy. An appropriate cutoff satisfying the covariant entropy bound can be chosen so that the entropy has a definite bound. Among the entropy for each of cosmological eras, the vacuum energy-dominated era turns out to give the maximal entropy which is in fact compatible with assumptions from the brick wall method. ", "introduction": " ", "conclusions": "" }, "0808/0808.3207_arXiv.txt": { "abstract": "% The dark clouds in the constellation of Chamaeleon have distances of 160-180~pc from the Sun and a total mass of $\\sim$5000~$M_\\odot$. The three main clouds, Cha~I, II, and III, have angular sizes of a few square degrees and maximum extinctions of $A_V\\sim5$-10. Most of the star formation in these clouds is occurring in Cha~I, with the remainder in Cha~II. The current census of Cha~I contains 237 known members, 33 of which have spectral types indicative of brown dwarfs ($>$M6). Approximately 50 members of Cha~II have been identified, including a few brown dwarfs. When interpreted with the evolutionary models of Chabrier and Baraffe, the H-R diagram for Cha~I exhibits a median age of $\\sim$2~Myr, making it coeval with IC~348 and slightly older than Taurus ($\\sim$1~Myr). The IMF of Cha~I reaches a maximum at a mass of 0.1-0.15~$M_\\odot$, and thus closely resembles the IMFs in IC~348 and the Orion Nebula Cluster. The disk fraction in Cha~I is roughly constant at $\\sim50$\\% from 0.01 to 0.3~$M_\\odot$ and increases to $\\sim65$\\% at higher masses. In comparison, IC~348 has a similar disk fraction at low masses but a much lower disk fraction at $M\\ga1$~$M_\\odot$, indicating that solar-type stars have longer disk lifetimes in Cha~I. ", "introduction": "The southern constellation of Chamaeleon contains one of the nearest groups of dark clouds to the Sun ($d\\sim160$-180~pc). An extinction map of these clouds is shown in Figure~\\ref{fig:map1} \\citep{dob05}. The main clouds in Chamaeleon have angular sizes of a few square degrees and are referred to as Chamaeleon I, II, and III \\citep{hof62}.\\footnote{In most publications, including this review, the designations Cha~II and Cha~III have been reversed from the original ones assigned by \\citet{hof62}.} Digitized Sky Survey (DSS) images of Cha~I and II are shown in Figure~\\ref{fig:map2} and a color optical image of Cha~I obtained by G.\\ Rhemann is shown in Figure~\\ref{fig:map3}. These clouds contain signposts of recent star formation in the form of several reflection nebulae, including Ced~110, 111, \\citep{ced46} and the Infrared Nebula \\citep[IRN,][]{sh83}. An optical image of the area surrounding Ced~111 is shown in Figure~\\ref{fig:ced111}. Newborn stars were first directly discovered in Chamaeleon through their variability and H$\\alpha$ emission \\citep{hof62,hen63,men72}. The masses of the Chamaeleon clouds and the stellar densities of young stars within them are low compared to many other star-forming regions. Because Chamaeleon is nearby and well-isolated from other young stellar populations, it has been a popular target for studies of low-mass star formation. \\begin{figure} \\includegraphics[width=\\textwidth]{f1.eps} \\caption{Extinction map of the Chamaeleon dark clouds \\citep{dob05}. The maximum extinction in this map is $A_V\\sim$10.} \\label{fig:map1} \\end{figure} \\begin{figure} \\includegraphics[width=\\textwidth]{f2.eps} \\caption{ DSS images of Cha~I ($2\\hbox{$^\\circ$}\\times2\\hbox{$^\\circ$}$) and Cha~II ($1\\fdg5\\times1\\fdg5$). The reflection nebulae Ced~111 and Ced~112 are associated with the B stars HD~97048 and HD~97300, respectively. } \\label{fig:map2} \\end{figure} \\begin{figure} \\includegraphics[width=\\textwidth]{f3.eps} \\caption{ A wide-field optical color-composite image of the Cha~I cloud ($1\\fdg4\\times2\\hbox{$^\\circ$}$). North is up and east is left. Courtesy G. Rhemann. } \\label{fig:map3} \\end{figure} \\begin{figure} \\includegraphics[scale=1.08]{f4.eps} \\caption{ An optical color-composite image of the Ced~111 reflection nebula in Cha~I obtained with VLT ($6\\farcm8\\times11\\farcm2$). Four of the brightest young stars within this area are labeled. North is up and east is left. Courtesy ESO. } \\label{fig:ced111} \\end{figure} ", "conclusions": "" }, "0808/0808.0132.txt": { "abstract": "The GHASP survey (Gassendi HAlpha survey of SPirals) consists of 3D \\ha~data cubes for 203 spiral and irregular galaxies, covering a large range in morphological types and absolute magnitudes, for kinematics analysis. It is the largest sample of Fabry-Perot data published up to now. In order to provide an homogenous sample, reduced and analyzed using the same procedure, we present in this paper the new reduction and analysis for a set of 97 galaxies already published in previous papers but now using the new data reduction procedure adopted for the whole sample. The GHASP survey is now achieved and the whole sample is reduced using adaptive binning techniques based on Voronoi tessellations. We have derived \\ha~data cubes from which are computed \\ha~maps, radial \\VFs~as well as residual \\VFs, \\PVMs, \\RCs~and kinematical parameters for almost all galaxies. The \\RCs, the kinematical parameters and their uncertainties are computed homogeneously using the new method based on the power spectrum of the residual \\VF. This paper provides the kinematical parameters for the whole sample. For the first time, the integrated \\Ha~profiles have been computed and are presented for the whole sample. The total \\Ha~fluxes deduced from these profiles have been used in order to provide a flux calibration for the 203 GHASP galaxies. This paper confirms the conclusions already drawn from half the sample concerning (i) the increased accuracy of \\PAs~measurements using kinematical data, (ii) the difficulty to have robust determinations of both morphological and kinematical inclinations in particular for low inclination galaxies and (iii) the very good agreement between the \\TF~relationship derived from our data and previous determinations found in the literature. ", "introduction": "\\label{intro} The GHASP survey consists of a large sample of spiral and irregular galaxies observed with a scanning Fabry-Perot for studying their kinematical and dynamical properties through the ionized hydrogen component. The goals of this study have been described in \\citet{Epinat:2008}. % The GHASP sample is by now the largest homogenous sample of Fabry-Perot data ever published, comprising 3D data for 203 galaxies. % This paper is the seventh of a series called hereafter Paper I to VI \\citep{Garrido:2002,Garrido:2003,Garrido:2004,Garrido:2005,Spano:2008, Epinat:2008} presenting the data obtained in the frame of the GHASP survey. The observations were lead during fourteen observing runs at the ``Observatoire de Haute Provence (OHP)'', France, from 1998 to 2004. The survey is now achieved. % The observing runs number 8 to 14 have been presented in Paper VI, which relies on a set of 108 galaxies, providing 106 \\VFs~and 93 \\RCs. Those data have been reduced with the new data reduction procedure (see Paper VI and references therein). % The data presented in this paper are those of the seven first observing runs (already presented in Papers I to IV) that have been re-reduced using the same method as in Paper VI. It also contains data for UGC 3382 and UGC 11300 that have been improved by adding new data (runs 3, 5 and 6) to the ones already published in Paper VI (runs 10 and 13). Thus, the data presented here consist of a set of 97 galaxies, providing 96 \\VFs~and 82 \\RCs. To be clear on the goals and limits of this present work, we summarize hereafter what we present and what we do not in this paper. We present: \\begin{itemize} \\item the new individual maps and \\PVMs~in Appendix \\ref{maps} (on line data only), \\item the new \\rcs, in Appendix \\ref{rc} and the new corresponding tables in Appendix \\ref{rc_tables}. \\end{itemize} In this paper, we lead the same kind of analysis as the one presented in Paper VI concerning: \\begin{itemize} \\item the study of the parameters of the kinematical models, \\item the study of the residual \\VFs, \\item the \\TF~relation. \\end{itemize} Because it is useful to display and analyze all the data together, we put here in the same tables (in Appendix \\ref{tables}) the new parameters and the results already published in Paper VI so that the reader does not have to compile tables coming from different publications. With respect to Papers I to IV, some distances and absolute magnitudes have been recomputed (using better estimations). %Furthermore, all the parameters for the whole GHASP sample are in %a single table and the reader does not have to compile tables %coming from different publications. %Thus we provide: %\\begin{itemize} % \\item all the parameters of the kinematical models (in section \\ref{xxx} and \\ref{tables}). % \\item all the residual \\VFs~(in section \\ref{xxx}), % \\item the \\TF~relation. %\\end{itemize} For the whole GHASP sample we make a new analysis on an absolute flux calibration made using the data calibrated by \\citet{James:2004} and the integrated \\ha~profiles deduced from our data cubes. Because this has been discussed in previous papers, we do not present any more: \\begin{itemize} \\item the morphological types and luminosity distributions of the whole GHASP sample (see Paper VI), \\item the data reduction procedure used here, including the computation of the \\RCs, the determination of the kinematical parameters and the determination of the uncertainties (see Paper VI), \\item the instrumental set-up of the instrument for the data re-reduced in this paper (see Papers I to IV), % \\item the log of the observation (see Papers I to IV) xxxpourtant on remet des info recapitulatives en table B1: que dire donc?xxx, \\item the individual comments for each galaxy (see Papers I to IV), except when the new reduction procedure leads to new comments or to conclusions noticeably different from the previous ones (see Appendix \\ref{notes}). \\end{itemize} In section \\ref{calibration} we make an indirect flux calibration of the \\ha~profiles. In section \\ref{analysis} we present the data and in section \\ref{tullyfisher} we compute the \\TF~relation. In section \\ref{conclusion} we give the summary and conclusions. When the distances of the galaxies are not known, a Hubble constant H$_{0}$=75\\kms~Mpc$^{-1}$ is used throughout this paper. ", "conclusions": "\\label{conclusion} The knowledge of the links between the kinematical and dynamical state of galaxies helps us to increase our understanding of the physics and evolution of galaxies. The GHASP sample, which consists of 203 spiral and irregular galaxies, covering a wide range of morphological types and absolute magnitudes, has been constituted in order to provide a kinematical reference sample of nearby galaxies. The galaxies have been observed in the \\ha~line using Fabry-Perot techniques, leading to the construction of data cubes. This sample is by now the largest set of galaxies ever homogeneously observed with Fabry-Perot techniques. Major improvements in the reduction (adaptive binning techniques, ghost suppression, treatment of faint outskirts regions, etc) and in the analysis (determination of the \\RC~and of the kinematical parameters and their uncertainties, etc) have been developed and implemented in the data reduction procedure and homogeneously applied to the whole GHASP sample (see Paper VI for additional details). In this paper, 97 galaxies have been re-reduced using adaptive binning techniques in order to provide homogeneous data for the whole sample. For each galaxy, we have presented the \\ha~\\VF, the \\ha~monochromatic image and eventually the \\ha~residual \\VF, the \\PVM~along the major axis and the \\RC, when available, leading for the whole sample to 200 \\VFs~and 177 \\RCs. From the data cubes, integrated \\Ha~profiles have also been produced. A post calibration has allowed to compute indirect absolute \\Ha~flux for all the galaxies belonging to the GHASP sample. This post calibration has been done using fluxes for 69 galaxies found in the literature \\citep{James:2004}. We confirm and strengthen most of the results already obtained from half the sample: \\begin{itemize} \\item A high quality model has been achieved to represent the axi-symmetric rotational component of the galaxies since no typical signatures for biases are observed in the residual \\VFs. This means that the residuals observed in the residual \\VF~are due to actual non circular motions and not to an uncorrect determination of the kinematical parameters (position of the center, \\PA, inclination and rotation velocity). In addition, the \\PVMs~confirm the validity of the \\RCs. \\item The mean residual velocity dispersion is strongly correlated with the maximum amplitude of the \\VF. For a given velocity range, this correlation does not clearly depend on the morphological type. However strongly barred galaxies have a higher residual velocity dispersion than mild-barred or unbarred galaxies. Peculiar galaxies also show a high residual velocity dispersion. \\item The determinations of kinematical \\pas~are robust whatever the inclination of the galaxy whereas morphological \\pas~are poorly determined for low inclination systems. Moreover, morphological \\PAs~have systematically higher uncertainties than kinematical ones. This is a major argument for deriving \\rcs~from integral field spectroscopy rather than long slit spectroscopy instruments that could lead to incorrect positioning of the slit (a difference between the morphological and kinematical \\PAs~larger than 30\\degr~is found for $\\sim$15$\\%$ of the GHASP sample). This may strongly bias mass distribution models and \\TF~studies. In order to build a mass model, the stellar mass distribution derived from the surface brightness profile is combined with the \\rc~deduced from the \\vf. The \\pas~of the major axis deduced from the surface brightness image and from the \\vf~should be identical. Important inconsistencies may appear if these \\PAs~are misaligned. \\item Galaxies with poor determination of their morphological \\pas~have usually unreliable and overestimated morphological inclinations. The agreement between kinematical and morphological inclinations is better when assuming a thin disk in particular for high inclination galaxies. For galaxies with intermediate disk inclinations (higher than 25\\degr~and lower than 75\\degr), to improve the quality of the \\RC, it is possible to reduce the degrees of freedom in kinematical models by fixing the inclination to the morphological value. This is specially true when only low quality kinematical data are available as it is the case for high redshift galaxies. \\item The use of the whole GHASP sample leads to a \\TF~relationship in perfect agreement with \\citet{Tully:2000}, despite important differences in the selection of both samples. With respect to the result presented in Paper VI, the use of the whole sample increases the agreement with \\citet{Tully:2000}. Three comments should be underlined: (i) galaxies with inclination lower than 25\\degr~are inappropriate for \\TF~relation determination since their estimated velocities are easily overestimated; (ii) fast rotators (V$_{max}>$300\\kms) are maybe less luminous (than expected from the \\TF~relation); (iii) for fast rotators and high luminosity galaxies, the agreement with the \\TF~relation is better when the morphological inclination of the galaxy is computed without taking into account the increasing thickness of the disk when the morphological type of the galaxies moves from early to late types. \\end{itemize} From these data and analysis, it is now possible to adress the scientific drivers on the whole GHASP sample in forthcoming papers." }, "0808/0808.1728_arXiv.txt": { "abstract": "\\noindent We present comprehensive models for the Herbig Ae stars MWC275 and AB~Aur that aim to explain their spectral energy distribution (from UV to millimeter) and long baseline interferometry (from near-infrared to millimeter) simultaneously. Data from the literature, combined with new mid-infrared (MIR) interferometry from the Keck Segment Tilting Experiment, are modeled using an axisymmetric Monte Carlo radiative transfer code. Models in which most of the near-infrared (NIR) emission arises from a dust rim fail to fit the NIR spectral energy distribution (SED) and sub-milli-arcsecond NIR CHARA interferometry. Following recent work, we include an additional gas emission component with similar size scale to the dust rim, inside the sublimation radius, to fit the NIR SED and long-baseline NIR interferometry on MWC275 and AB~Aur. In the absence of shielding of star light by gas, we show that the gas-dust transition region in these YSOs will have to contain highly refractory dust, sublimating at $\\sim$1850K. Despite having nearly identical structure in the thermal NIR, the outer disks of MWC275 and AB~Aur differ substantially. In contrast to the AB~Aur disk, MWC275 lacks small grains in the disk atmosphere capable of producing significant 10-20 $\\mu$m emission beyond $\\sim$7AU, forcing the outer regions into the ``shadow'' of the inner disk. ", "introduction": "\\label{intro} Herbig Ae (HAe) stars are pre-main-sequence stars of intermediate mass (1.5-3 solar masses). They exhibit a robust excess in emission over stellar photospheric values from near-infrared (NIR) to the millimeter (mm) wavelengths. This excess is now attributed to the passive reprocessing of stellar light by dust in the circumstellar environment \\citep{lkha2001, natta2001, dullemond2001}. The geometry of the circumstellar environment of HAe stars has been actively debated in the astronomy community over the last two decades. Some of the early workers in this field \\citep{Hillen92} showed that spectral energy distribution (SED) of HAe stars could be explained by emission from circumstellar matter in disk-like geometry. Others \\citep{Mirosh97} argued that the emission could also arise from dust in a spherical geometry around the star, proving the inadequacy of SED modeling alone in uniquely fixing the geometry of the circumstellar matter. The first observational evidence in favor of a disk geometry came from millimeter (mm) interferometry in the form of asymmetries detected \\citep{Man97} in the mm images. Asymmetries in the NIR emission were also detected by the Palomar Test-Bed Interferometer \\citep{eisner2003, eisner2004}, settling the debate in support of a disk geometry for circumstellar material in Herbig Ae stars. Most interferometric studies of HAe stars have relied on simple geometric models \\citep{Man97, rmg1999, rmg2001, eisner2003, eisner2004, monnier2005} that explain the emission geometry of the system in only narrow wavelength ranges. This method, albeit extremely useful in elucidating some of the morphology details, is not adequate for exploring the interdependency in structure of the inner and outer parts of the disk. A number of studies \\citep{dullemond2001, dullemond2004, boekel2005b} have shown that the structure of the inner disk at fractions of an AU scale clearly affects the structure of the outer disk. A complete understanding of the circumstellar disk structure in HAe stars therefore requires models that simultaneously explain the SED and interferometry over a large wavelength range. Such models have begun to appear in the literature only recently \\citep{ponto2007, kraus2007}. In this paper, we develop comprehensive disk models to explain the SED and interferometry of the HAe stars MWC275 and AB Aur. MWC275 and AB~Aur are prototype pre-main-sequence stars of similar ages and spectral type with extensive circumstellar disks. Due to the availability of photometric and interferometric data over a large wavelength range, MWC275 and AB~Aur are ideal candidates for testing disk models for YSOs. The extent of their circumstellar-dust disks was first measured by \\citet{Man97} to be several 100AU using the Owens Valley Radio Observatory (OWRO). \\citet{Natta2004} resolved the MWC275 disk in the mm and reported a de-convolved, projected dust-disk size of 300AU$\\times$180AU. More recently, \\citet{isella2007} analyzed IRAM, SMA and VLA continuum and $^{12}$CO, $^{13}$COand $^{18}$CO line data constraining the gas-disk radius to be 540AU with the gas in Keplerian rotation around the central star. Scattered light studies of MWC275 \\citep{grady2000} and AB~Aur \\citep{grady1999, oppenheimer2008} show the presence of arcs and rings in the circumstellar disk. \\citet{corder2005} resolved the AB~Aur CO disk radius to be $\\sim$600AU, finding strong evidence for Keplerian rotation for the bulk of the disk. \\citet{corder2005} and \\citet{lin2006} detected spiral arms in CO emission with radii of $\\sim$150AU, while \\citet{fukagawa2004} detected similar structure in Subaru H-band scattered light images. AB~Aur also has substantial envelope material on scales larger then 600 AU \\citep{grady1999, semenov2005, corder2005, lin2006}. MIR emission probes the giant planet formation region in circumstellar disks \\citep{calvet1992, cg97, dullemond2001} with the emission arising from warm dust (T $>$ 150K). \\citet{Meeus2001} and \\citet{boekel2005} used the 10$\\mu$m MIR silicate emission feature from MWC275 and AB~Aur to show that dust grains in these systems had grown larger than the typical interstellar medium grain sizes. \\citet{marinas2006} imaged AB~Aur at 11.7$\\mu$m and found the emission FWHM size to be 17$\\pm$4 AU consistent with the flared disk models of \\citet{dullemond2004}. In this paper, we present new 10$\\mu$m measurements of AB~Aur and MWC275 with the Keck Segment Tilting Experiment \\citep[described in \\S\\ref{data}]{Monnier2004}. In contrast to AB~Aur, the MWC275 disk is unresolved by the Segment Tilting Experiment (maximum baseline of 10m), requiring the VLT Interferometer (100m baseline) to probe to its MIR structure \\citep{Leinert2004}. These observations suggest that MWC275 disk differs considerably from AB~Aur and we present a detailed comparison of the two disk structures in the discussion (\\S\\ref{discuss}). Thermal NIR emission probes hot regions (typically the inner AU) of the disk with temperatures greater than 700K. The NIR disks of MWC275 and AB~Aur were first resolved with IOTA by \\cite{rmg1999, rmg2001} and subsequently observed at higher resolution with PTI \\citep{eisner2004}, Keck Interferometer \\citep{monnier2005} and the CHARA interferometer array \\citep{Tannirk2008}. In \\citet[hereafter T08]{Tannirk2008} we showed that inner-disk models in which majority of the K-band emission arises in a dust rim \\citep{dullemond2001, isella2005, Tannirk2007} fail to fit the CHARA data at milli-arcsecond resolution. We also demonstrated that the presence of additional NIR emission (presumably from hot gas) inside the dust destruction radius can help explain the CHARA data and the NIR SED. First calculations for the effects of gas on rim structure \\citep{muzerolle2004} showed that for plausible disk parameters, presence of gas does not modify dust-rim geometry significantly. Besides a poorly understood interferometric visibility profile, MWC275 also displays as yet ill-understood NIR and MIR SED time variability \\citep{sitko2007} which has been interpreted as variations of the inner disk structure. In \\S\\ref{MWC275_model} and \\S\\ref{ABAur_model} we present a detailed analysis of the NIR visibility and SED for MWC275 and AB~Aur, placing constraints on the wavelength dependence of the opacity source inside the dust destruction radius. In this study, we focus on (i) explaining the inner-disk structure and discuss important open problems and (ii) modeling the MIR emission morphology of the disks and the shape of the MIR spectrum. The paper is organized into 7 sections with \\S\\ref{data} detailing the observations. \\S\\ref{model} explains the disk model and the modeling strategy. \\S\\ref{MWC275_model} and \\S\\ref{ABAur_model} analyze MWC275 and AB~Aur SED and visibilities in relation to the disk models. We present a discussion on our results and our conclusions in \\S\\ref{discuss} and \\S\\ref{conclusion} respectively. ", "conclusions": "\\label{conclusion} We have presented the first set of comprehensive disk models for the SED and interferometry of Herbig Ae stars MWC275 and AB~Aur. We have shown that `standard' models for the dust evaporation front where the bulk of the near-infrared emission arises from a dust wall, fail to explain the near-infrared spectral energy distribution and interferometry. Standard models produce large bounces in visibility at high spatial frequency, which is not observed in the data. We have conclusively demonstrated that the presence of an additional smooth emission component (presumably hot gas) inside the dust destruction radius and on a similar size scale to the dust rim can ameliorate the situation. In the absence of shielding of star light by gas, we have established that dust grains in the gas-dust transition region will have to be highly refractory, sublimating at ~1850K. The small mid-infrared size of MWC275 relative to AB~Aur, shows that the dust grains in the outer disk MWC275 are significantly more evolved/settled than the grains in the AB~Aur disk. We suggest that dynamical processes (like planetesimal collisions) that maintain the population of micron-sized grains producing the 10$\\mu$m feature in the spectrum, are operational only in the inner ~7AU of MWC275. However, in AB-Aur the small-dust producing mechanisms exist at least out to ~20 AU and maybe even beyond." }, "0808/0808.1172_arXiv.txt": { "abstract": "{}{The high-redshift ($z=4.048)$ gamma-ray burst GRB 060206 showed unusual behavior, with a significant rebrightening by a factor of $\\sim4$ at about 3000 s after the burst. We argue that this rebrightening implies that the central engine became active again after the main burst produced by the first ejecta, then drove another more collimated jet-like ejecta with a larger viewing angle. The two ejecta both interacted with the ambient medium, giving rise to forward shocks that propagated into the ambient medium and reverse shocks that penetrated into the ejecta. The total emission was a combination of the emissions from the reverse- and forward- shocked regions. We discuss how this combined emission accounts for the observed rebrightening.} {We apply numerical models to calculate the light curves from the shocked regions, which include a forward shock originating in the first ejecta and a forward-reverse shock for the second ejecta.} {We find evidence that the central engine became active again 2000 s after the main burst. The combined emission produced by interactions of these two ejecta with the ambient medium can describe the properties of the afterglow of this burst. We argue that the rapid rise in brightness at $\\sim3000$ s in the afterglow is due to the off-axis emission from the second ejecta. The precession of the torus or accretion disk of the central engine is a natural explanation for the departure of the second ejecta from the line of sight.}{} ", "introduction": "The gamma-ray burst (GRB) 060206 at Galactic Coordinates $l=78.07$ deg, $b=78.28$ deg triggered $Swift$-\\textbf{BAT} on February 6th, 04:46:53 UT (trigger time $t=0$) (Morris et al. 2006). It exhibited a single peak, with a duration of ${T_{90}=7\\pm 2}$ s and a total fluence of $8.4\\pm0.4\\times10^{-7}$ $\\rm{erg/cm^2}$ in the 15-350 keV band (Palmer et al. 2006). The spectroscopic redshift $z$ is 4.048 (Fynbo et al. 2006). Applying the peak energy ${E_{\\rm{peak}}}={75.4\\pm19.5}$ keV, the best-fit low energy photon index ${\\Gamma_1=1.06\\pm 0.34}$ and a fixed high energy photon index $\\Gamma_2=2.5$, the isotropic-equivalent energy integrated from 1 to $10^4$ keV in the explosion rest frame is ${E_{\\gamma,\\rm{iso}}=5.8\\times10^{52}}$ erg (Palmer et al. 2006). $Swift$-$\\rm{XRT}$ began to observe this burst 58 s after the \\textbf{BAT} trigger time. At the same time, $Swift$-UVOT started the on-target monitoring and detected the optical afterglow (Boyd et al. 2006). A number of ground-based telescopes performed follow up observations. The 2-m robotic Liverpool Telescope began to observe it at $t=309$ s and carried out multicolor $r'i'z'$ photometry. In the R-band the light-curve exhibited three obvious bumps in the first 75 minutes including a steep rise ($\\Delta r'\\approx -1.6$ at $t\\approx 3000$ s) (Monfardini et al. 2006). About 48.1 minutes later after the trigger time, the Rapid Telescopes for Optical Response (RAPTOR) system at Los Alamos National Laboratory began to take optical images. The obtained light curve confirmed the rebrightening from ${r'\\sim 17.3}$ to a peak value $r'\\sim 16.4$. The subsequent decline to $r'{\\sim 16.75}$ at $t=80$ min was followed by a secondary rebrightening by $\\Delta r'\\sim -0.1$ around $t=90$ min ($\\rm{Wo\\acute{z}niak}$ et al. 2006). The MDM telescope observed a smooth break at $t_b=0.6$ days with another bump at $t\\approx16000$ s. The overall X-ray light curve has a similar shape as the optical light curve (Stanek et al. 2007). One of the most remarkable features of this burst is that the optical light curve had a significant rebrightening and exhibited small ``bumps'' and ``wiggles''. Similar bumps and wiggles have also been seen in a number of optical afterglows (Stanek et al. 2007). GRB 970508 had an optical afterglow light curve rather similar to that of GRB 060206 (Galama et al. 1998). The optical light curve of another recent burst, GRB 060210, also displayed a rebrightening at time $t\\sim500$ s and a shallow decay in the early epoch. The above ``unusual'' behavior , which is not predicted by the standard fireball model, may be more the norm than the exception (Stanek et al. 2007). Possible scenarios for the remarkable rebrightening in GRB 060206 at $\\sim$3000 s are a renewed energy injection (Rees \\& \\Mesz 1998; Kumar \\& Piran 2000; Sari \\& \\Mesz 2000) and a density-jump in the circum-burst medium (Dai \\& Lu 2002). However, as discussed by Monfardini et al. (2006), the X-ray band frequency is above the cooling frequency at ${t\\sim 3000}$ s, so the flux does not depend on the ambient density (Freedman \\& Waxman 2001). Nakar \\& Granot (2007) showed that even a sharp and large increase in the ambient medium density cannot produce a significant rebrightening as seen in the afterglow. So the rebrightening cannot be due to a density jump in the ambient medium. If the rebrightening is caused by energy injection, a huge impulsive energy injection ${\\Delta E\\sim 1.8 E_0}$ at $\\sim$3000 s was required, where $E_0$ is the blast wave energy before the rebrightening. In this paper, we present an alternative scenario. The central engine of this burst become active again after the initial burst and ejects another more collimated jet with a larger viewing angle. This jet and the initial jet sweep up the interstellar medium (ISM). The multi-wavelength emission predicted by this model can reproduce both the observed X-ray and optical data. The observational results are presented in Sect 2. We describe the scenario in Sect 3 and fit the remarkable optical rebrightening of GRB 060206 in Sect 4. Finally, we summarize our results and discuss their implications in Sect 5. ", "conclusions": "We have presented a solution for the remarkable rebrightening observed in the afterglow of GRB 060206, which is attributed to emission from an off-axis beamed jet originating from the late activity of the central engine after the prompt gamma-ray emission phase. We have argued that the precession of the torus or accretion disk of the central engine has caused the two jets to move in different directions. Although we only attempt to describe the large rebrightening, it is reasonable to speculate that the five bumps in the R-band afterglow light curve of similar profiles may have the similar origins: emission from different delayed off-axis jets ejected by the GRB central engine. The difference between the temporal indices of the five post-bump light curves indicates that the electrons in different jets have different spectral indices $p$ for their electron distributions. As a result, the spectra exhibit an SED evolution, especially when the separate jets have comparable contributions to the total emission. This can be naturally explained by the significant SED evolution in GRB 060206 detected by the Liverpool Telescope (Monfardini et al. 2006). The isotropic energy of the second jet is more than one order of magnitude higher than the first jet in our scenario. The collimation-corrected energy of Jet $B$ is $\\sim 4.3$ times larger than Jet $A$, which is larger than the energy required to explain the big rebrightening in the energy injection model as mentioned in Sect 1. The precise mechanism that triggers the central engine again remains unknown. One possibility is that a mass of debris falls back onto the central compact object, generating another more energetic jet. Since our scenario can reproduce well the observations of the large bump in GRB 060206, as a conservative extrapolation, we propose that GRB 970508 and GRB 060210, which display remarkable rebrightening, may have in addition a precessing torus or accretion disk. A further consequences of our scenario is that the jet break is determined by the off-axis second jet rather than the first one, which produces the main burst. In this case, we cannot measure the isotropic energy of the second jet directly and must fit the afterglow data to obtain an estimation. When a large bump appears in GRB afterglows, we are therefore unlikely to be able to derive the jet opening angle, using the break time and isotropic gamma-ray energy release, because these two quantities originate in two different jets (Stanek et al. 2007). Finally, several models could be applied to explain afterglow light curves exhibiting rebrightening. These include variable external density profiles (Lazzati et al. 2002), refreshed shocks (Granot et al. 2003; \\Bj et al. 2004), and angular dependence of the energy profile on the jet structure (Nakar et al. 2003), each of which can play a role. Peculiar behavior in light curves caused by the precession of the central engine was discussed by Reynoso et al. (2006), although more observations are required to identify its true nature." }, "0808/0808.2492_arXiv.txt": { "abstract": "{ The deceleration mechanisms for relativistic jets in active galactic nuclei remain an open question, and in this paper we propose a model which could explain sudden jet deceleration, invoking density discontinuities. This is particularly motivated by recent indications from HYbrid MOrphology Radio Sources, suggesting that in some case Fanaroff-Riley classification is induced by variations in density of the external medium. } {Exploiting high resolution, numerical simulations, we demonstrate that for both high and low energy jets, always at high Lorentz factor, a transition to a higher density environment can cause a significant fraction of the directed jet energy to be lost on reflection. This can explain how one-sided jet deceleration and a transition to FR I type can occur in HYbrid MOrphology Radio Sources, which start as FR II (and remain so on the other side).} {For that purpose, we implemented in the relativistic hydrodynamic grid-adaptive AMRVAC code, the Synge-type equation of state introduced in the general polytropic case by Meliani et al. (2004). To demonstrate its accuracy, we set up various test problems in appendix, which we compare to exact solutions that we calculate as well. We use the code to analyse the deceleration of jets in FR II/FR I radio galaxies, following them at high resolution across several hundreds of jet beam radii. } { We present results for 10 model computations, varying the inlet Lorentz factor from 10 to 20, including uniform or decreasing density profiles, and allowing for cylindrical versus conical jet models. As long as the jet propagates through uniform media, we find that the density contrast sets most of the propagation characteristics, fully consistent with previous modeling efforts. When the jet runs into a denser medium, we find a clear distinction in the decelaration of high energy jets depending on the encountered density jump. For fairly high density contrast, the jet becomes destabilised and compressed, decelerates strongly (up to subrelativistic speeds) and can form knots. If the density contrast is too weak, the high energy jets continue with FR II characteristics. The trend is similar for the low energy jet models, which start as underdense jets from the outset, and decelerate by entrainment in the lower region as well. We point out differences that are found between cylindrical and conical jet models, together with dynamical details like the Richtmyer-Meshkov instabilities developing at the original contact interface. } {} ", "introduction": "Accretion disks surrounding black holes, jets found in association with compact objects, and Gamma Ray Bursts (GRBs) all represent violent astrophysical phenomena. They are associated with relativistic flows, both with respect to the occuring velocities and to their prevailing equation of state. The jets from Active Galactic Nuclei (AGN) and in GRBs are accelerated in a short distance to reach a high Lorentz factor: typical values are $\\gamma\\sim 5-30$ \\citep{Kellermannetal04} for AGNs and $\\gamma>100$ for GRBs \\citep{Woods&Loeb95, Sari&Piran95}. In this acceleration phase, situated at the base of the jet, it is believed that jet energy is dominated by thermal energy and Poynting flux, and that a fraction of these energies contributes to the efficient acceleration of the flow. Thereby, the relativistic fluid changes its state from relativistic, corresponding to an effective polytropic index $\\Gamma=4/3$, to classical (polytropic index $\\Gamma=5/3$) when the thermal energy is converted to kinetic energy. Also further in the jet paths, where the jets interact with the surrounding medium, a fraction of the directed kinetic energy is converted to thermal energy at the shock fronts formed. According to the prevailing Lorentz factor of the jet, these shocks could be relativistic and therefore very efficient to convert kinetic to thermal energy, or could be Newtonian and only weakly efficient in compression. To investigate the occuring relativistic flows, a realistic equation of state must therefore be adopted, to handle both classical as well as relativistic temperature variations in space and time. For that purpose, we implemented in the relativistic (magneto)hydrodynamic grid-adaptive AMRVAC code~\\citep{Melianietal07, Keppens03, vanderHolst07}, the Synge-type equation of state introduced for a general polytropic case as in Meliani et al. (2004) and as applied in the adiabatic case by Mignone \\& McKinney (2007). To demonstrate its accuracy, we set up various stringent test problems in an appendix, which we compare to exact solutions for Riemann problems, that we calculate as well. The latter include Riemann problems at Lorentz factors of order $\\gamma\\approx 100$, which represent extreme values relevant for Gamma Ray Burst flows. The development of relativistic numerical hydrodynamic codes can help us understand the physics of astrophysical jet propagation. In the last decade, significant progress was made in numerical special relativistic hydrodynamic (HD) and magnetohydrodynamic codes. Various authors worked on the development of conservative shock-capturing schemes which use either exact or approximate Riemann solver based methods for relativistic hydrodynamics \\citep{Eulderink&Melemma94, Fontetal94, Aloyetal99, DelZanna&Bucciantini02, Mignone&Bodo05} (for a review see \\cite{Marti&Muller03}). The study of relativistic hydrodynamic fluids benefits also from using spatial and temporal adaptive techniques, or Adaptive Mesh Refinement \\citep{Duncan&Hughes94, Zhang&MacFadyen06, Melianietal07, Wangetal07}. The various numerical simulations usually involve a simplified equation of state (EOS) with a constant polytropic index. Notable exceptions with more realistic EOS treatments are found in \\cite{Mignone05, Mignone&McKinney07, Perucho&Marti07}. We present in the appendix to this paper the required extension of the AMRVAC code \\citep{Melianietal07, Keppens03, vanderHolst07} with the more realistic polytropic EOS introduced by~\\cite{Melianietal04} which is based on the Synge gas equation, the relativistically correct perfect gas law~\\citep{Synge57, Mathews71}. The equations which we solve, and the schemes used, are also mentioned there. Many previous investigations of relativistic jet propagation through the interstellar medium (ISM) in radio galaxies concentrate on uniform ISM conditions \\citep{Duncan&Hughes94, Martietal97, KomissarovetFall98, Aloyetal99}. These investigations contributed to our understanding of the jet deceleration, where one then distinguishes dynamics for Faranoff-Riley type I (FR I) and FR II radio galaxies, according to the power of the jet and hence the accretion rate in their galactic center. In the FR I radio galaxies, the associated jets are relativistic on parsec scale and sub-relativistic on kiloparsec-scales, so that jet deceleration, and thus energy redistributions, must happen on kiloparsec scales \\citep{Hardcastleet05}. Various studies have looked into possible deceleration processes with both non-relativistic and relativistic HD codes. \\cite{Hooda&wiita96} and \\cite{Hooda&wiita98} investigated with a classical HD code the propagation of a 3D jet through an interface between dense ISM and less dense intercluster medium. They found that for such interface marking a density decrease, the jet does not decelerate when crossing it, and that the deceleration mostly happens in the lower region (dense ISM). Moreover, in their study the jets are not deflected at the interface ISM/ICM. \\cite{Normanetal88} also use a classical code to study the sudden deceleration of jets, when the jet crosses a shock wave in the external medium. Other works investigate the interaction of jets with dense clouds in both non-relativistic hydrodynamic simulations \\citep{Saxtonetal05} and relativistic simulations \\citep{Choietal07}. \\citet{Duncan&Hughes96} study relativistic jet propagation in uniform overdense media and use an equation of state for a pair-plasma. \\citet{Schecketal02} study the influence of the matter composition of a high jet energy jet on its interaction with the external medium. They investigate the two extreme cases of pure leptonic and baryonic plasmas. \\citet{Perucho&Marti07} examined the propagation of the low energy jet of the specific FR I radio source 3C 31 using an elaborate hydrodynamics model, with an equation of state that distinguishes the contribution of leptons and baryons to density and pressure. Most recently, \\citet{Rossietal08} investigate the propagation of jets in uniform overdense media in 3D, using a realistic synge EOS. \\begin{table*} \\caption{The most relevant characteristics and parameters for various selected relativistic hydrodynamic simulations in the literature: DH94 \\citep{Duncan&Hughes94}, M97 \\citep{Martietal97}, H98 \\citep{Hardeeetal98}, R99~\\citep{Rosenetal99} S02~\\citep{Schecketal02}, PM07 \\citep{Perucho&Marti07}, R08 \\citep{Rossietal08}. Our simulations are indicated by MKG. We mention the type of the simulation 2D or 3D and use of AMR, jet beam kinetic luminosity $L_{\\rm b}$ (when available), $\\gamma_{\\rm b}$ Lorentz factor of the beam, $M_{\\rm b}$ relativistic Mach number, $\\eta$ density ratio (``d\" means density variation), $\\theta$ open angle of the jet, $R_{\\rm b}/r_{\\rm cell}$ grid cells through jet beam, type of EOS (\"Lep\" means Leptonic, \"Bar\" Baryonic), endtime of the simulation in units of light crossing time of the jet beam radius.} \\label{table:1} % \\centering % \\begin{tabular}{c c c c c c c c c c} \\hline\\hline % Paper & case & $L_{\\rm b}$ & $\\gamma_{\\rm b}$ & $M_{\\rm b}$ & $\\rho_{\\rm medium}/\\rho_{\\rm jet}$& $\\theta_{\\rm b} $& $R_{\\rm b}/r_{\\rm cell} $ &EOS& Size in $R_{\\rm b}$ (and time)\\\\ % \\hline % \\hline % DH94/R99 & A (2D, AMR) & - &1.048 & 6 & $10.0$&0& 24& $5/3$& $10 \\times 41.67 $ (-)\\\\ DH94/R99 & B (2D, AMR) & - &5.0 & 8 & $10.0$&0& 24& $5/3$& $16.67 \\times 41.67$ (-)\\\\ DH94/R99 & C (2D, AMR) & - &10.0 & 8 & $10.0$ &0& 24& $5/3$& $16.67\\times 41.67 $ (-)\\\\ DH94/R99 & D (2D, AMR) & - &10.0 & 15 & $10.0$ &0& 24& $4/3$& $16.67 \\times 41.67$ (-)\\\\ M97 & A1 (2D) & - &7.1 & 9.97 & $10^2$ &0& 20& $4/3$& $10.5 \\times 50$ (50)\\\\ M97 & A2 (2D) & - &22.37 & 31.86 & $10^2$ &0& 20& $4/3$& $10.5 \\times 50$ (50)\\\\ M97 & a1 (2D) & - &7.1 & 9.97 & $1.0$ &0& 20& $4/3$& $10.5 \\times 25$ (25)\\\\ M97 & a2 (2D) & - &7.1 & 9.97 & $10.0$ &0& 20& $4/3$& $10.5 \\times 25$ (25)\\\\ M97 & B1 (2D) & - &7.1 & 41.95 & $10^2$ &0& 20& $4/3$& $10.5 \\times 50$ (50)\\\\ M97 & B2 (2D) & - &22.37 & 132.32 & $10^2$ &0& 20& $4/3$& $10.5 \\times 50$ (50)\\\\ M97 & b1 (2D) & - &2.29 & 13.61 & $10.0$ &0& 20& $4/3$& $10.5 \\times 25$ (25)\\\\ M97 & b2 (2D) & - &7.1 & 41.95 & $10.0$ &0& 20& $4/3$& $10.5 \\times 25$ (25)\\\\ M97 & C1 (2D) & - &2.29 & 13.61 & $10^2$ &0& 20& $5/3$& $10.5 \\times 50$ (50)\\\\ M97 & C2 (2D) & - &7.1 & 41.95 & $10^2$ &0& 20& $5/3$& $10.5 \\times 50$ (50)\\\\ M97 & C3 (2D) & - &22.37 & 132.32 & $10^2$ &0& 20& $5/3$& $10.5 \\times 50$ (50)\\\\ M97 & c1 (2D) & - &2.29 & 13.61 & $10.0$ &0& 20& $5/3$& $10.5 \\times 25$ (25)\\\\ M97 & c2 (2D) & - &7.1 & 41.95 & $10.0$ &0& 20& $5/3$& $10.5 \\times 25$ (25)\\\\ H98 & B (2D, AMR) & - &5.52 & 8.87 & $10.0$ &0& 24& $5/3$& $16.67 \\times 41.67$ (140.0)\\\\ H98 & C (2D, AMR) & - &14.35 & 11.52 & $10.0$ &0& 24& $5/3$& $16.67 \\times 41.67$ (21.43)\\\\ H98 & D (2D, AMR) & - &10.0 & 15.0 & $10.0$ &0& 24& $4/3$& $16.67 \\times 41.67$ (35.273)\\\\ H98 & E (2D, AMR) & - &2.55 & 8.53 & $10.0$ &0& 24& $5/3$& $16.67 \\times 41.67$ (140.0)\\\\ S02 & A (2D) & $10^{46}$ &6.62 & 9.24 & $10^5$ &0& 6& Lep& $200\\times 500$ (5950)\\\\ S02 & B (2D) & $10^{46}$ &6.62 & 14.33 & $10^3$ &0& 6& Lep& $200\\times 500$ (5400)\\\\ S02 & C (2D) & $10^{46}$ &7.95 & 130.14 & $10^3$ &0& 6& Bar& $200\\times 500$ (5300)\\\\ PM07 & A (2D) & $10^{44}$ &2.0 & 4.75 & $10^5$ ``d\" &0& 16& Lep/Bar& $200\\times 450$ (37231)\\\\ R08& A (3D) & - &10.0 & 28.3 & $10^2$ &0& 20& Synge& $50 \\times 150 \\times 50$ (240)\\\\ R08& B (3D) & - &10.0 & 28.3 & $10^4$ &0& 20& Synge& $60 \\times 75 \\times 60$ (760)\\\\ R08& C (3D) & - &10.0 & 28.3 & $10^4$ &0& 12& Synge& $50 \\times 75\\times 50$ (760)\\\\ R08& D (3D) & - &10.0 & 300 & $10^4$ &0& 20& Synge& $50 \\times 150\\times 50$ (760)\\\\ R08& E (3D) & - &10.0 & 300 & $10^2$ &0& 12& Synge& $24 \\times 200\\times 24$ (150)\\\\ MKG & A (2D, AMR) & $10^{46}$ &20.0 & 1200 & $0.1496$-$4.687$ &0& 64& Synge& $10 \\times 400$ (380)\\\\ MKG & B (2D, AMR) & $10^{46}$ &20.0 & 1200 & $0.1496$-$671.22 $ &0& 128& Synge& $10 \\times 400$ (900)\\\\ MKG & C (2D, AMR) & $10^{43}$ &10.0 & 39 & $36.52$-$203.66$ ``d\" &0& 76& Synge& $10 \\times 400$ (820)\\\\ MKG & D (2D, AMR) & $10^{43}$ &10.0 & 39 & $36.52$-$148.15$ ``d\" &1& 76& Synge& $10 \\times 400$ (820)\\\\ MKG & E (2D, AMR) & $10^{43}$ &20.0 & 35 & $146$-$847.46$ ``d\" &0& 76& Synge& $40 \\times 400$ (820)\\\\ MKG & F (2D, AMR) & $10^{43}$ &20.0 & 35 & $146$-$645.16$ ``d\" &1& 76& Synge& $40 \\times 400$ (820)\\\\ MKG & G (2D, AMR) & $10^{46}$ &10.0 & 1300 & $0.033$-$0.495$ ``d\" &0& 144& Synge& $40 \\times 400$ (380)\\\\ MKG & H (2D, AMR) & $10^{46}$ &10.0 & 1300 & $0.033$-$30.6748$ ``d\" &1& 76& Synge& $40 \\times 400$ (800)\\\\ MKG & I (2D, AMR) & $10^{46}$ &10.0 & 1300 & $0.033$-$0.306748$ ``d\" &1& 76& Synge& $40 \\times 400$ (480)\\\\ MKG & J (2D, AMR) & $10^{46}$ &20.0 & 1200 & $0.1496$-$12.95$ ``d\" &1& 76& Synge& $40 \\times 400$ (480)\\\\ \\hline % \\end{tabular} \\end{table*} In this paper, we investigate a scenario of relativistic jet propagation through an ISM with a sudden jump in density. Since jets propagate over enormous distances, it is inevitable that they encounter density jumps in the external medium. As we assume that matter in the inner part of the radio galaxy has been cleared away during the evolution of the radio galaxy, we typically follow jet dynamics through a lighter medium at first, which then suddenly changes far away to a denser medium. In a sequence of model runs, we vary the jet and medium properties to cover both high and low jet beam kinetic luminosities, straight as well as conical jets (varying the jet opening angle), and allow for either uniform or decreasing density profiles. To put our work into perspective and to appreciate its relation to previous works better, we compiled in Table~\\ref{table:1} the most relevant parameters and simulation specifics for selected relativistic hydro models. The most novel assumption is the transition from low to high density across a contact interface (across which the pressure is necessarily constant), which is different from the work done by \\citet{Hooda&wiita98} where they use a classical code with a density decrease across the contact, and also different from the study by \\citet{Lokenetal93}, where they use a classical code to look into jet propagation through a pressure wall. We explore the influence of a sudden density jump in ISM on jet propagation, stability, and formation of the bow shock. We aim to better understand the jet efficiency in transporting energy and mass from the central AGN regions to the denser ISM at large distances, where the jet cocoon gets formed. Note also from Table~\\ref{table:1} that our models extend the previous works most notably by typically covering much larger distances than previously studied (as we need to obtain representative endstate morphologies in both lower and upper media), and are invariably in the high Lorentz factor regime, combined with a high resolution through the jet beam. The high energy jets we consider start out as denser than their immediate surroundings (a likely property deduced from all common jet launch scenarios, but previously ignored by all jet simulations to date). Moreover, the density contrasts we investigate between jet and outer medium are much more reasonable than the 5 orders of magnitude density differences previously used by~\\citet{Schecketal02}. In what follows, we first motivate our model assumptions and the simulation setup in Section~\\ref{secmodel}. We discuss our main findings in Section~\\ref{secres}. ", "conclusions": "In this paper, we extended, presented and applied the relativistic hydrodynamics AMRVAC code \\citep{Melianietal07,vanderHolst07} with a generalized variable polytropic index equation of state for the purpose of modeling the relativistic shocks in GRBs and AGN jets. As follows in the appendix, we used various shock-capturing schemes on a variety of test cases for adiabatic and non-adiabatic cases for code validation. We demonstrated the code ability with stringent new test problems, developed according to the astrophysical context. We tested the code also for a case with a heated flow, using the Synge-like equation of state from \\cite{Melianietal04}. The Riemann problems were also solved exactly, and the AMR simulations handled cases with Lorentz factors of order 100 accurately. We explored a new scenario for sudden deceleration of relativistic jets in the FR I radio galaxies. This model for jet deceleration is based on a density jump in the external medium, with density suddenly increasing to the upper medium. We include models of conical jets with an initial opening angle, and allow for decreasing denisty profiles within the upper medium. We investigated the propagation and the dynamics of relativistic AGN jets over a long distance, resolving small scale instabilities which develop in the cocoon and which could be responsible for particle acceleration. This study was only possible with adaptive mesh refinement. We quantified and discussed the deceleration of the jet resulting from its interaction with a layered medium. We point out that jet deceleration can be significantly aided by an internal oblique shock that forms in jets with conical expansion. Under fairly extreme (but not as extreme as pursued in other simulations to date) density conditions in the dense upper region, it can reproduce the strong deceleration observed in FR I jets. In the early phase, the head of the cold fast (beam Lorentz factor 10 or 20) jet propagates in the low density medium at a high Lorentz factor $\\gamma=5$. In cases with high energy, almost no cocoon or backflows develop in this phase, and as the jet is heavier than the external region, it behaves more ballistically. When observed in this region, the relativistic beaming is high, so that one would observe a one-sided jet. In low energy jets, a cocoon always forms and slow backflows develop. They in turn induce multiple internal shocks making the jet decelerate and re-accelerate. In jets with low energy and with an initial conical flow, the flow remains ballistic until it reaches a self-consistently forming oblique shock. This compresses and decelerates the jet and two regions appear in the jet beam, the first with high relativistic beaming, and the second with low Lorentz factor, showing clear disturbances by the surrounding cocoon. In the outer region where the density increases suddenly, we find increased entrainment of ambient material through velocity shear instabilities at the jet boundary and working surfaces. Note that we here for the first time address how the prior interaction of the jet with the low density ambient medium in the inner region (which makes the head of the jet consist of swept-up ambient medium and a shocked beam) affects jet propagation in denser media. The pre-structured jet head has lower Mach number, which gives rise to a strong interaction with the denser ambient medium. For models with a weak increase in density in the external medium, the cocoon shows little turbulent structure and no backflow appears. The jet in this case remains stable and relativistic in the denser upper region. Also a fraction of its energy is deposited in the external medium, but only through the frontal shock. This model could be relevant for jets in FR II. With this scenario, one can explain how most of the energy of the jet is deposited in the outer region, since relatively little jet energy gets transferred to the medium in the inner region. Those models with more extreme increases in density in the external medium, showed the formation of an overpressured and turbulent cocoon. A strong backflow develops in the outer region, which disturbs the jet structure. The backflow compresses the jet and induces internal shocks which give rise to knot formation. The jet becomes subrelativistic within the simulated domain. The models with initial conical jets show the development of an oblique shock that decelerates the jet. In future work, we will follow these jets over increasingly larger distances in 3D scenarios. We also intend to provide synthetic observational radio maps from our computed jet models. \\appendix\\label{appendixA}" }, "0808/0808.2812_arXiv.txt": { "abstract": "We report the discovery by the Robotic Optical Transient Search Experiment (ROTSE-IIIb) telescope of SN 2008es, an overluminous supernova (SN) at $z=0.205$ with a peak visual magnitude of $-$22.2. We present multiwavelength follow-up observations with the \\textsl{Swift} satellite and several ground-based optical telescopes. The ROTSE-IIIb observations constrain the time of explosion to be 23$\\pm$1 rest-frame days before maximum. The linear decay of the optical light curve, and the combination of a symmetric, broad H$\\alpha$ emission line profile with broad P Cygni H$\\beta$ and Na I~$\\lambda5892$ profiles, are properties reminiscent of the bright Type II-L SNe 1979C and 1980K, although SN 2008es is greater than 10 times more luminous. The host galaxy is undetected in pre-supernova Sloan Digital Sky Survey images, and similar to Type II-L SN 2005ap (the most luminous SN ever observed), the host is most likely a dwarf galaxy with $M_{r} > -17$. \\textsl{Swift} Ultraviolet/Optical Telescope observations in combination with Palomar 60 inch photometry measure the spectral energy distribution of the SN from 200 to 800 nm to be a blackbody that cools from 14,000 K at the time of the optical peak to 6400 K 65 days later. The inferred blackbody radius is in good agreement with the radius expected for the expansion speed measured from the broad lines (10,000 km s$^{-1}$). The bolometric luminosity at the optical peak is $2.8 \\times 10^{44}$ erg s$^{-1}$, with a total energy radiated over the next 65 days of $5.6 \\times 10^{50}$ erg. The exceptional luminosity of SN 2008es requires an efficient conversion of kinetic energy produced from the core-collapse explosion into radiation. We favor a model in which the large peak luminosity is a consequence of the core-collapse of a progenitor star with a low-mass extended hydrogen envelope and a stellar wind with a density close to the upper limit on the mass-loss rate measured from the lack of an X-ray detection by the \\textsl{Swift} X-Ray Telescope. ", "introduction": "} Hydrogen-rich supernovae (SNe II) produce their radiative energy in several phases; with a diversity of luminosities, light curves, and spectroscopic features that divide them into subclasses. An initial burst of UV/X-ray radiation is observed at the time of shock breakout (Schawinski \\etal 2008; Gezari \\etal 2008a), followed by declining UV/optical emission from the adiabatic expansion and cooling of the SN ejecta (e.g., SN II-pec 1987A, Hamuy \\etal 1988; SN IIb 1993J, Schmidt \\etal 1993; Richmond \\etal 1994, SN II-P 2006bp, Immler \\etal 2007). Type II-plateau (II-P) SNe are characterized by their subsequent plateau in optical brightness, which is understood as the result of a progenitor with a substantial hydrogen envelope, for which a cooling wave of hydrogen recombination recedes through the inner layers of the ejecta (Falk \\& Arnett 1977; Litvinova \\& Nadyozhin 1983). Type II-linear (II-L) SNe do not have a plateau phase, but rather a steep drop in luminosity that is thought to be the result of the ejection of large amounts of radioactive material (Young \\& Branch 1989) or a small hydrogen envelope mass (Barbon \\etal 1979; Blinnikov \\& Bartunov 1993). The subclass of bright SNe II-L ($M_{B} < -18$; Patat \\etal 1994) has been proposed to be the result of an extended low-mass envelope (Swartz \\etal 1991) and reprocessing of UV photons in the superwind of the progenitor star (Blinnikov \\& Bartunov 1993) or a gamma-ray burst (GRB) explosion buried in a hydrogen-rich envelope (Young \\etal 2005). Type II-narrow (IIn) SNe show signs of interaction of the ejecta with circumstellar material (CSM) in the form of narrow emission lines, X-ray emission, and an excess in optical luminosity (e.g., SN 1988Z: Turatto \\etal 1993). After the photospheric phase in SNe II, a final exponential decline is powered by heating due to the radioactive decay of $^{56}$Co. In this paper, we present the discovery of an overluminous SN II that has the properties of an SN II-L, but with an exceptional luminosity that we argue is best explained by the core-collapse of a progenitor star with a low-mass, extended hydrogen envelope and a dense stellar wind. We disfavor more exotic scenarios such as an interaction with a massive shell of CSM expelled in an episodic eruption, a pair-instability explosion, or a buried GRB. In \\S \\ref{sect_obs}, we present the Robotic Optical Transient Experiment (ROTSE-IIIb) discovery data, and follow-up observations with the \\textsl{Swift} satellite and several ground-based telescopes; in \\S \\ref{sec_disc} we compare the properties of SN 2008es to other known Type II SNe, and use our observations to constrain the mechanism powering the extraordinary radiative output; and in \\S \\ref{sec_conc} we make our conclusions. ", "conclusions": "\\label{sec_conc} The overluminous SN 2008es ($M_{V}=-22.2)$ is the second most luminous SN observed, and in the same league as the extreme SNe 2005ap, 2006gy, and 2006tf, which had peak visual magnitudes of $-22.7$, $-22.0$, and $-20.7$, respectively. In Figure \\ref{fig_sne}, we compare the absolute $R$-band light curves of the 4 SNe (all discovered by the ROTSE-IIIb telescope) in rest-frame days since the time of explosion. SN 2006gy and 2006tf are classified spectroscopically as Type IIn SNe because of their narrow peaked P Cygni Balmer lines (Ofek \\etal 2007; Smith \\etal 2008), although with deviations from the smooth profiles typical of SNe IIn in SN 2006gy (Smith \\etal 2007). The time of explosion for SN 2005ap is estimated by Quimby \\etal (2007a) to be $\\sim 3$ weeks before maximum. The time of explosion for SN 2006tf is poorly constrained (Smith \\etal 2008), so for comparison we assume that it takes the same time to reach the peak as SN 2008es. Of the SNe in the same luminosity class, the light curve of SN 2008es is most like SN 2005ap, the most luminous SN ever observed \\citep{2007ApJ...668L..99Q}. Interestingly, SN 2005ap also has a dwarf galaxy host, with $M_{R} = -16.8$. However, SN 2005ap does differ from SN 2008es in the fact that it demonstrated absorption features at early times, as well as P Cygni absorption in its broad H$\\alpha$ profile. This is not consistent with the spectra of bright II-L SNe which have decreasing P Cygni absorption with increasing luminosity (Patat \\etal 1994). The extended envelope/wind model is problematic for SN 2005ap, since it would be hard to produce early absorption features without corresponding emission. The extreme luminosity and linear decay of SN 2005ap were attributed instead to a collision of the SN ejecta with a dense circumstellar shell, a GRB explosion buried in a H envelope, or a pair-instability eruption powered by radioactive decay \\citep{2007ApJ...668L..99Q}. The peak luminosity, linear decay, and spectroscopic features of SN 2008es place it in a subclass of ultra-bright Type II-L SNe. The light curve and SED of SN 2008es, measured in detail from the UV through the optical, point toward a core-collapse explosion of a nonstandard progenitor star with a super wind and extended envelope. Wide-field optical synoptic surveys such as the ROTSE-III Supernova Verification Project, Pan-STARRS, and LSST will continue to explore the large parameter space of core-collapse explosions, and increase our understanding of the effects of variations in progenitor star properties and environments. {\\bf Note Added in Proof:} More recent $V$-band observations from \\textsl{Swift}, as well as $V, R, I$ measurements from the MDM 2.4m, reveal that the slope of the light curve is steeper than the decay rate of $^{56}$Co up to $\\sim$145 rest-frame days after maximum. Therefore, radioactive decay must not yet be the dominant source of the luminosity from the SN, and the inferred upper limit on the initial $^{56}$Ni mass is securely within the range of normal Type II SNe (Hamuy 2003)." }, "0808/0808.0483_arXiv.txt": { "abstract": "We report on a problem found in {\\sc Mercury}, a hybrid symplectic integrator used for dynamical problems in Astronomy. The variable that keeps track of bodies' statuses is uninitialised, which can result in bodies disappearing from simulations in a non-physical manner. Some {\\sc fortran} compilers implicitly initialise variables, preventing simulations from having this problem. With others compilers, simulations with a suitably large maximum number of bodies parameter value are also unaffected. Otherwise, the problem manifests at the first event after the integrator is started, whether from scratch or continuing a previously stopped simulation. Although the problem does not manifest in some conditions, explicitly initialising the variable solves the problem in a permanent and unconditional manner. ", "introduction": "\\label{intro} {\\sc Mercury} \\citep{Chambers1999} is a general-purpose software package, written in {\\sc fortran77}, for doing N-body integrations to investigate dynamical problems in Astronomy. In a set of simulations to study accretion dynamics using {\\sc Mercury} \\citep{Torres2008}, it was found that the results contained fewer planets than expected. Analysis of output files and creation of scripts and movies to carefully follow the processes showed discontinuities in the number of embryos during the integrations. No events for these bodies were being registered in the output files and their disappearance was non-physical. ", "conclusions": "\\label{conclusions} We repeated all the tests showed in Section \\ref{charac} with a corrected version of {\\sc Mercury}, initialising the variable array {\\sc stat} at the end of the subroutine MIO\\_IN in the source code. No discontinuities were seen in any results, independently of conditions. The new version is now being used in 40 new simulations of accretion dynamics. We believe this small change can improve the program making it more reliable for any type of N-body problem simulations. The corrected version can be downloaded from \\url{http://www.astro.keele.ac.uk/~dra/mercury/}." }, "0808/0808.3449_arXiv.txt": { "abstract": "We test the proposal by El-Zant \\etal\\ that the dark matter density of halos could be reduced through dynamical friction acting on heavy baryonic clumps in the early stages of galaxy formation. Using $N$-body simulations, we confirm that the inner halo density cusp is flattened to $0.2$ of the halo break radius by the settling of a single clump of mass $\\ga 0.5\\%$ of the halo mass. We also find that an ensemble of 50 clumps each having masses $\\ga 0.2\\%$ can flatten the cusp to almost the halo break radius on a time scale of $\\sim9\\;$Gyr, for an NFW halo of concentration 15. We summarize some of the difficulties that need to be overcome if this mechanism is to resolve the apparent conflict between the observed inner densities of galaxy halos and the predictions of LCDM. ", "introduction": "\\label{intro} The LCDM model for structure formation in the universe has been very successful on large scales (\\eg\\ Springel, Frenk \\& White 2006), but has run into a number of difficulties on galaxy scales. One difficulty is that the predicted dark matter density in the halos of bright galaxies appears to be greater than that inferred from the best observational data (\\eg\\ Weiner \\etal\\ 2001; Dutton \\etal\\ 2007) or from dynamical friction constraints (Debattista \\& Sellwood 2000); see Sellwood (2008b) for a review. A number of possible solutions have been suggested. Binney, Gerhard \\& Silk (2001) and others suggest that the halo density can be reduced by feed-back from star formation, but Gnedin \\& Zhao (2002) show that the density cannot be reduced by more than a factor two, even for the most extreme feedback, unless the baryons are unreasonably concentrated. Weinberg \\& Katz (2002) propose that dynamical friction between a bar and the halo can reduce the central density, but Sellwood (2008a) shows that significant reductions require extreme bars and remove a large fraction of the angular momentum from the baryons. A third possible solution was proposed by El-Zant, Shlosman \\& Hoffman (2001, hereafter EZ01), who suggested that dynamical friction between heavy gas clumps and dark matter would transfer energy to the dark matter, thereby reducing its inner density. Essentially EZ01 propose a process of mass segregation that allows the baryonic matter to displace the dark matter. If it works, it would amount to baryon settling with halo {\\it de}-compression, as Dutton \\etal\\ (2007) argued would be required to improve agreement of LCDM predictions with scaling relations of $\\sim L_*$ galaxies. Kassin, de Jong \\& Weiner (2006) also remark that rotation curves could be more easily reconciled with LCDM predictions if halo compression could somehow be avoided. Tonini, Lapi \\& Salucci (2006) use the idea to predict rotation curves for galaxies. Mo \\& Mao (2004) apply the mechanism to pre-process small halos in the very early stages of structure formation. A similar, but significantly distinct, idea was tested by Gao \\etal\\ (2004), who argue that some unspecified attractor mechanism causes violent relaxation processes to reset the collisionless mass profile of the combined dark and baryonic matter. Since violent mergers disrupt disks, Gao \\etal\\ clearly invoke a different process from the slow frictional in-spiral of baryonic clumps proposed by EZ01. EZ01 suppose that the baryons collect into dense mass clumps which then sink to the center by dynamical friction. They assume: (1) the clumps are small enough they do not collide and (2) are sufficiently tightly bound that they are not tidally disrupted. They also implicitly assume that the dense clumps (3) maintain their coherence independent of any internal physical processes (such as possible star formation) and (4) they require the mass clumps to contain little dark matter. The in-spiral of massive clumps has been studied extensively and we do not attempt a complete list of references. Van den Bosch \\etal\\ (1999) estimate the infall rate of dark matter sub-clumps through dynamical friction. Ma \\& Boylan-Kolchin (2004) simulate a mass spectrum of satellites each composed of particles moving within a main halo; they find that the combined density profile of the inner halo and disrupted satellites can either steepen or flatten, depending on the properties of the clumps. Arena \\& Bertin (2007) study both the strength of the friction force and the effect on the density profile for both cuspy and cored initial galaxy profiles. Here we undertake a direct test of the proposal by EZ01, using self-consistent $N$-body simulations, in order to determine the clump mass required to effect a significant density reduction. We first derive the settling time-scale for single massive clumps in cuspy halos, and then present simulations containing an ensemble of heavy clumps in a background of light halo particles. We discuss the physical plausibility of their assumptions in the concluding section. El-Zant \\etal\\ (2004, hereafter EZ04) apply the same idea to account for the flattened density profile reported by Sand \\etal\\ (2002) in the galaxy cluster MS 2137-23. In this context, they identify the galaxies themselves as the heavy baryonic clumps that settle to the cluster center. Both they, and Nipoti \\etal\\ (2004), present simulations to show that the in-spiral of galaxies can cause the background dark matter density profile to flatten, although both studies neglect the DM halos attached to the infalling galaxies. In an appendix, we show that our simulations are in good agreement with those of EZ04 for the same problem. ", "conclusions": "El-Zant \\etal\\ (2001) proposed that dynamicl friction on massive gas clumps could lead to the outward displacement of dark matter. They estimated the magnitude of the effect on a galaxy halo using the Chandrasekhar dynamical friction formula. They followed up (El-Zant \\etal\\ 2004) with simulations of a galaxy cluster, in which the heavy particles were the individual galaxies, finding some density decrease. We present simulations of softened heavy particles moving under the influence of gravitational forces within a background NFW halo composed of a large number of light particles. We confirm that the inner halo density is lowered by dynamical friction on heavy particles, as has been found previously (\\eg\\ Ma \\& Boylan-Kolchin 2004; Arena \\& Bertin 2007), and show that larger reductions result from more massive particles that also settle more rapidly. We specifically test the model proposed by EZ01, who divided all the baryons into a number of equal mass clumps that were already somewhat concentrated towards the halo center. We find (Fig.~\\ref{collect}) that the $\\sim 10\\%$ baryonic mass fraction must be entirely made up of $\\la 150$ equal clumps if the cusp in the dark matter is to be flattened to $r \\la r_s$ within 300 dynamical times, which scales to $\\sim 9\\;$Gyr for a $c=15$ halo. The principal reason that we observe a more mild density reduction than predicted by EZ01 is that friction is significantly weaker than they assumed. They, and Tonini \\etal\\ (2006), adopted $\\ln\\Lambda \\simeq \\ln\\eta \\simeq 8.5$ whereas we find $\\ln\\Lambda \\simeq 2.5$ (\\S\\ref{tscale}), which causes the mass clumps to settle 2-3 times more slowly than they assumed, with a corresponding reduction in the rate energy is given up to the halo. More massive clumps both spiral in more rapidly and displace more dark matter (Fig.~\\ref{massch}). Ma \\& Boylan-Kolchin (2004) presented a pair of simulations that revealed a much smaller density change when the three most massive clumps were eliminated compared with the result when they were retained. Thus substantial reductions of the inner halo density require just a few extremely massive clumps, assuming they are not disrupted as they settle. The time scale, in years, varies as $\\rho_s^{-1/2}$, and therefore would be longer in lower concentration halos. At higher redshifts, the orbital period is a fixed fraction of the age of the universe at a constant overdensity (since both scale as the square root of the density in an Einstein-de Sitter universe). While halo densities rise with redshift, their relative over-densities are lower (\\eg\\ Zhao \\etal\\ 2003), and therefore more massive clumps are required to have any effect in the time available; Mo \\& Mao (2004) invoke baryon clumps having 5\\% of the halo mass when $2 \\la z \\la 5$. EZ04 argue that the same physical process applies to galaxies in clusters, and present supporting simulations. Again adopting their numerical model and assumptions, we confirm (Appendix) their result that the density cusp flattens for $r \\la 0.2r_s$, where we also demonstrate that the result is independent of the numerical method. The physical assumptions underlying these successes are more questionable, however. Purely baryonic gas clumps are indeed expected to form through the Jeans instability as gas cools and settles in the main halo (\\eg\\ Maller \\& Bullock 2004), but the high-resolution experiments of Kaufmann \\etal\\ (2006) show that gas fragments formed in this way have masses ranging up to $\\sim 10^6\\;$M$_\\odot$ only, some two orders of magnitude smaller than required to have interestingly short settling times. Dark matter sub-halos (\\eg\\ Diemand, Kuhlen \\& Madau 2007) may contain gas and would spiral in more rapidly. However, settling of such sub-halos would bring both baryons and DM into the center, diluting the desired separation of baryons from dark matter. The resulting changes to the overall dark matter density profile will clearly be less substantial. When Ma \\& Boylan-Kolchin (2004) treat sub-clumps as collections of particles, they find that the stripped mass is added to the main DM halo in such a way as approximately to replace the mass scattered to larger radii by dynamical friction. (The net effect can be either a small increase or decrease in the inner dark matter density, depending on the clump mass spectrum.) If dynamical friction is to separate the baryons from dark matter with the desired efficiency, baryonic mass clumps in sub-halos must somehow be stripped of their dark matter with high efficiency, without dissolving the gas clumps themselves. In the case of the cluster simulation, EZ04 describe the heavy particles as purely baryonic galaxies. Here again, the change in the central dark matter density will not be as substantial when the dark halos of the galaxies are taken into account, as also noted by Nipoti \\etal\\ (2004). Additional density reduction would be possible if massive baryonic clumps in the center of the halo could be re-accelerated. Models of this kind have been proposed by Mashchenko, Couchman \\& Wadsley (2006, 2007), who invoke star formation activity, and by Peirani, Kay \\& Silk (2008), who use AGN jets. It should be noted, however, that massive clumps of dense gas are not easily accelerated by these astrophysical processes (\\eg\\ MacLow \\& Ferrara 1999). The rapid density reduction reported by Mashchenko \\etal\\ (2006) occurs because they employed a gas clump of $10^8\\;{\\rm M}_\\odot$ within the cusp driven sinusoidally with a bulk speed that peaked at 18~km~s$^{-1}$. Most halo density reduction occurs as clumps plunge rapidly within the cusp (Fig.~\\ref{massfr}), as noted by EZ01. Therefore, clumps that start out in the cusp fall in more rapidly and are most efficient at displacing the dark matter. However, only a small fraction ($\\la 5\\%$) of the baryons start out in the cusp, if they are distributed as the dark matter. The desired halo density reduction would be more rapid if the baryonic fraction in the cusp could be increased, but it is hard to see how such an initial mass segregation could have arisen without compressing the halo. Thus the challenge, if the proposal by EZ01 is to help solve the central density problem of dark matter halos, is to find ways to collect gas into clumps exceeding $\\sim 0.5\\%$ of the halo mass that can settle as coherent entities into the center." }, "0808/0808.3480_arXiv.txt": { "abstract": "The dynamics of small global perturbations in the form of linear combination of a finite number of non-axisymmetric eigenmodes is studied in two-dimensional approximation. The background flow is assumed to be an axisymmetric perfect fluid with the adiabatic index $\\gamma=5/3$ rotating with power law angular velocity distribution $\\Omega \\propto r^{-q}$, $1.5>1$. The stability of toroidal flows was studied extensively in the 1980s and 1990s by means of the traditional eigenvalue analysis which has been widely used since the well-known investigations of Lord Rayleigh. The growing non-axisymmetric {\\it definite frequency modes} were discovered in the barotropic tori with a positive radial gradient of the specific angular momentum. This phenomenon was cold the Papaloizou-Pringle instability in honour of the authors who published their work in 1984. In numerous subsequent papers this sort of instability was studied in two- and three-dimensional approximations by both analytical and numerical methods \\cite{b24,b27,b1,b25,b3,b26,b28,b2,b22}. It was found that there are growing surface gravity and sound waves. The general picture of the instability is constructed by these modes which grow due to the mutual coupling as well as by the Landau mechanism, i.e. due to interaction with the background flow \\cite{b8,b5}. However, as the rotation profile approaches the Keplerian law the gravity surface waves become damping while the sonic instability ceases because of very small increments \\cite{b3,b4}. This fact actually dropped the subject since the near-Keplerian rotation is of the most interest in connection with the problem of angular momentum transport in the accretion disk theory. Besides, the growing linear modes can exist owing to additional degrees of freedom in the flow. For instance, \\cite{b39} and \\cite{b40} claimed that stratorotational instability emerges in vertically stratified rotationally stable flows in the shearing sheet approximation. However, instability of axisymmetric flows can be considered from a quite different viewpoint. It is closely associated with the formulation of the {\\it initial value problem} for perturbations. Indeed, to examine the stability thoroughly one must show that {\\it any} initial perturbation will not ever affect the background flow. However, the absence of growing eigenmodes guarantees only the asymptotic stability and tells us nothing about the behaviour of perturbations at finite time intervals. The generalized approach to the hydrodynamical instability was realized and declared in the most distinctive form among the fluid physicists in the middle 1990s \\cite{b9,b10}. It is often called the nonmodal stability theory and concentrates on the initial value problem, partially on searching for optimal initial perturbations that can exhibit large transient growth even in asymptotically stable flows. Yet the role of this linear mechanism is usually emphasized in connection with the transition to turbulence \\cite{b32}. According to the so called bypass concept the linear dynamics is responsible for extraction of the perturbation energy from the background. It is supposed that nonlinear processes only redistribute energy between the modes. The general description and some references to hydrodynamical literature on this subject can also be found in \\cite{b13} and in application to astrophysical problems in the more recent paper \\cite{b14}. Mathematically, the evolution of small perturbations is governed by a linear dynamical operator which in general can be normal or non-normal. In the first case eigenmodes are orthogonal and the eigenanalysis gives the complete picture of the temporal evolution of perturbations. Strictly speaking, the largest possible growth of perturbations is determined at any moment by the highest eigenvalue \\cite{b21}. In the case of non-normal dynamical operator eigenmodes become non-orthogonal and the substantial temporal growth of perturbations is possible even if there are no growing eigenmodes \\cite{b9}. The nonmodal approach to {\\it local} perturbations was first pointed out also in astrophysical literature in \\cite{b11} and \\cite{b12}. The local framework allows us to perform the spatial Fourier transform and thus to study directly the temporal behaviour of perturbations. This direction was developed then in a number of papers \\cite{b18,b31,b30,b29,b15,b16,b17}. Another branch of astrophysical investigations that use nonmodal approach is devoted to global dynamics in turbulent accretion disks. For instance, \\cite{b19} considered the dynamics of small stochastic perturbations in an incompressible Keplerian disk and found the resulting steady-state coherent structures enhancing the angular momentum transfer outward. It is also worth mentioning the work \\cite{b20} which studied the transient growth of axisymmetric perturbations in a thin accretion disk with account of compressibility and vertical structure of the flow. Here we concentrate on the global two-dimensional dynamics of the bounded laminar axisymmetric flows which were studied actively 15-20 years ago. We do not introduce viscosity since its action is negligible over the first $\\sim 10-50$ rotation periods but we include compressibility which is necessary to do since the sound speed is always comparable or less than the shear velocity difference when we are dealing with free boundaries. First we briefly describe the solution of the boundary problem, i.e. the calculation of eigenmodes. This will be followed by the discussion of the optimization algorithm which allows us to determine the specific group of eigenmodes with the strongest transient growth. This is actually the standard mathematical procedure that has been used by many authors. Then we present the results and make the conclusions. ", "conclusions": "Our study shows the significance of the transient growth concept in application to astrophysical problems. We considered here the dynamics of {\\it global} perturbations in a sense that free boundary conditions which are natural in astrophysical practice were involved. Previous studies of a pure hydrodynamical instability in axisymetric flows with free boundaries have been carried out by methods of the traditional eigenanalysis, i.e. by searching for growing modes with exponential dependence on time. This way was proved to be unable to solve the problem of angular momentum transfer in almost Keplerian flows with finite compressibility, since the surface gravity eigenmodes are damping \\cite{b3} and the growing sonic eigenmodes have negligible increments. However, the present two-dimensional analysis demonstrates that situation noticeably changes if one turns to the dynamics of a linear {\\it combination} of eigenmodes. In fact, we studied mainly a finite number of slow neutral eigenmodes from acoustic spectrum of the basic flow with $\\omega8\\,\\sigma$) in the range $0^{\\circ}-34^{\\circ}$. Nonetheless, the optical colors are independent of inclination below $\\approx 12^{\\circ}$, showing no evidence for a break at the reported boundary between the so-called dynamically ``hot'' and ``cold'' populations near $\\approx 5^{\\circ}$. The commonly accepted parity between the dynamically cold CKBOs and the red CKBOs is observationally unsubstantiated, since the group of red CKBOs extends to higher inclinations. Our data suggest, however, the existence of a different color break. We find that the functional form of the color-inclination relation is most satisfactorily described by a non-linear and stepwise behavior with a color break at $\\approx 12^{\\circ}$. Objects with inclinations $\\geqslant 12^{\\circ}$ show bluish colors which are either weakly correlated with inclination or are simply homogeneously blue, whereas objects with inclinations $<12^{\\circ}$ are homogeneously red. ", "introduction": "The Kuiper Belt is a disk of icy bodies having semi-major axes larger than that of Neptune. Its members are usually known as Kuiper Belt Objects (KBOs) or Trans-Neptunian Objects (TNOs). The distribution of their orbits is structured leading to the identification of several dynamical families. Resonant KBOs are those which are trapped in mean-motion resonances with Neptune (those trapped in the 3:2 resonance are also known as Plutinos). Scattered KBOs, also known as Scattered Disk Objects, are essentially highly eccentric KBOs under strong gravitational influence of Neptune. Classical KBOs (CKBOs) possess relatively circular orbits that are neither located in any strong mean-motion resonance with Neptune nor strongly subject to its gravitational influence. Since their discovery in 1992 more than 1200 KBOs have been identified. Due to their faintness only about 50 can be spectroscopically studied with the currently available instruments. Multicolor photometry provides, however, a first-order approximation of their spectra, hence of their surface composition. Most KBOs can be studied photometrically and about 230 objects have at least one measured color. Their surface colors have shown to be most diverse, ranging from neutral and even slightly blue (relative to the Sun) to extremely red, suggesting a large compositional diversity \\citep[see review by][]{Dor+08}. The origin of the color diversity remains unclear. Various suggestions have been made in the context of collisional resurfacing \\citep{LuuJew96, GilH02, Del+04}. Nevertheless, none of the proposed models has been able to consistently explain the colors \\citep{JewLuu01, TheDor03, Del+04}. Another possibility is that the observed color differences reflect primordial compositional variations \\citep[e.g.][]{TRC03}. Such compositional differences would be hard to explain if KBOs formed {\\it in situ}, since the temperature difference between 30 and 50 AU is a very modest $\\approx 10$ K. However, larger temperature and compositional differences might be possible if the KBO population, or part of it, did not form in place. Some dynamical models, in fact, suggest outward migration of KBOs \\citep[e.g.][]{Malho95, Gomes03}. Although there are no detailed chemical studies to address how varied such compositions would be and how these would reflect in the surface colors, these dynamical models imply a link between the current orbital inclinations of classical KBOs and their presumed location of origin. As a whole, KBOs do not show significant correlations between their colors and orbital parameters such as semi-major axis or perihelion distance. On the other hand, the CKBOs do show a correlation between orbital inclination and optical color \\citep{TegRom00, TruBro02} and a correlation between perihelion and color \\citep{TegRom00, Peix+04}. In parallel, several works have pointed out the existence of two groups of CKBOs with a separation at $\\approx 5^{\\circ}$ in inclination. The two groups, usually referred to as ``cold'' ($i\\lesssim 5^{\\circ}$) and ``hot'' ($i\\gtrsim 5^{\\circ}$), have been identified using not only orbital properties \\citep{Brown01, Ell+05} but also physical properties such as size and binarity \\citep{LS01,Peix+04,Gul+06, Noll+08}. Given its potential importance, we re-examine the color-inclination relation using a new data set and appropriate statistical tests. ", "conclusions": "The main goal of this work has been to investigate the color-inclination trend seen for Classical Kuiper Belt Objects (CKBOs). We have analyzed a sample of $B-R$ colors of 69 objects, excluding 2 apparent outliers as well as colors with error bars larger than $0.21$. Objects were classified as CKBOs according to a definition by \\cite{LykMuk07}. Orbital inclinations, denoted $i_k$, were calculated relative to the Kuiper Belt Plane following \\cite{Ell+05}. Our results may be summarized as: \\begin{enumerate} \\item The linear $B-R$ color-inclination correlation of CKBOs measured over the full range of inclinations from $0\\degr$ to $34\\degr$ is $\\rho=-0.70^{+0.09}_{-0.07}$, corresponding to a significance level larger than $8\\,\\sigma$. This is a strong and highly significant correlation, consistent with previously published values. \\item In contrast, the $B-R$ colors of CKBOs with inclinations $i_k\\leqslant 12.0^{\\circ}$$^{+0.5}_{-1.5}$ are statistically uncorrelated with inclination and are well described by $B-R=1.71\\pm0.11$. \\item CKBOs with $i_k >12.0^{\\circ}$$^{+0.5}_{-1.5}$ show a slight color vs. inclination dependence following $(B-R)=-0.0159\\,i_k+1.703$. The data are also formally consistent with a constant but bluer color, $B-R = 1.33\\pm0.20$, for $i_k \\geqslant 12.5\\degr$, and a constant red color $B-R=1.70\\pm0.11$ for $i_k <12.5\\degr$. \\item The data provide no evidence for a break or change in the $B-R$ color distribution at the boundary between the dynamically hot and cold populations, purportedly near $i_k \\approx 5^{\\circ}$. In this sense, we find no observational support for the frequently-cited parity between red CKBOs and the dynamically cold population. The CKBOs are red up to $i_k \\approx 12^{\\circ}$ and, therefore equally red into the dynamically hot population. \\end{enumerate}" }, "0808/0808.1811_arXiv.txt": { "abstract": "~~~\\\\ \\begin{center} {\\bf Abstract} \\end{center} The detection of a ``Cold Spot\" in the CMB sky could be explained by the presence of an anomalously large spherical underdense region (with radius of a few hundreds ${\\rm Mpc}/h$) located between us and the Last Scattering Surface. Modeling such an underdensity with an LTB metric, we investigate whether it could produce significant signals on the CMB power spectrum and bispectrum, via the Rees-Sciama effect. We find that this leads to a bump on the power spectrum, that corresponds to an ${\\cal O}(5\\%-25\\%)$ correction at multipoles $5 \\leq \\ell \\leq 50$; in the cosmological fits, this would modify the $\\chi^2$ by an amount of order unity. We also find that the signal should be visible in the bispectrum coefficients with a signal-to-noise $S/N\\simeq {\\cal O} (1-10)$, localized at $10 \\leq \\ell \\leq 40$. Such a signal would lead to an overestimation of the primordial $f_{NL}$ by an amount $\\Delta f_{NL}\\simeq 1$ for WMAP and by $\\Delta f_{NL}\\simeq 0.1$ for Planck. ", "introduction": "The recent WMAP~\\cite{WMAP} experiment has measured with great accuracy the anisotropies of the Cosmic Microwave Background (CMB), whose features are in good agreement with the expectations from inflation of a Gaussian spectrum of adiabatic fluctuations, fully described by a nearly scale-invariant power spectrum. However, it has been pointed out by several authors that the data contain various unexpected features. Some of them are localized at very large angular scales, such as the low quadrupole, the alignment of the low multipoles (for a review see~\\cite{reviewlowl} and references therein) and the power asymmetry between the northern and southern hemispheres~\\cite{hemispherical}. Another anomaly is the presence of the so-called Cold Spot~\\cite{ColdSpot1,ColdSpot2,ColdSpot3}, which is a large circular region on an angular scale of about $10^\\circ$ that appears to be anomalously cold: the probability that such a pattern would appear from Gaussian primordial fluctuations is estimated to be about $1.8\\%$~\\cite{ColdSpot1,ColdSpot2,ColdSpot3}. So, while this could still be a statistical fluke, some authors have put forward the idea that it could instead be due to an anomalously large spherical underdense region of some unknown origin, on the line of sight between us and the Last Scattering Surface (LSS)~\\cite{Tomita,InoueSilk}. We may also remind that~\\cite{rudnick} has claimed that, looking at the direction of the Cold Spot in the Extragalactic Radio Sources (the NVSS survey), an underdense region is visible at redshift $z\\sim 1$ (see, however~\\cite{huterer} for a paper that challenges this claim). Another motivation for studying such objects is that an underdense region of two or three hundreds of ${\\rm Mpc}/h$ could be enough to give an acceptable fit to the Supernova data and the CMB without Dark Energy~\\cite{ABNV}, if we happen to live near its centre. In this paper we also take the point of view that the Cold Spot could be due to such an underdense region, customarily denoted as a ``Void''. By modeling it through an inhomogeneous Lema\\^itre-Tolman-Bondi (LTB) metric (which also requires an overdense compensating shell), we compute the impact of such a Void on some observational quantities, focusing on the statistical properties of the CMB: in particular on the two-point correlation function (power spectrum) and the three-point correlation function (the bispectrum). This is interesting for the following reasons. First, if there is such a Void, does the power spectrum get a sizable correction? And, if yes, to what extent the estimation of the cosmological parameters is affected? Second, if there is such a Void, could it be detectable also in the bispectrum? Third, does the presence of such an object interfere with the measurement of a primordial non-gaussianity, which constitutes a very important piece of information in order to distinguish between models of inflation? It is well known that, passing through a Void, photons suffer some blue-shift due to the fact that the gravitational potential is not exactly constant in time -- the so-called Rees-Sciama (RS) effect~\\cite{ReesSciama}. By RS effect we mean the one associated to the variation of the gravitational potential from non-linear effects. There can be an effect already at the linear level, usually referred to as the Integrated Sachs-Wolfe (ISW) effect, which however would vanish in a matter dominated flat Universe. The ISW effect would be significant only when a Dark Energy component becomes dominant with respect to matter. Hence, here we focus our attention only on the RS effect, which is always present, and briefly comment on the extension to a $\\Lambda$CDM Universe later on. Note that there is a second physical effect due to a Void on the line of sight, in addition to the RS blue-shift of the photons: the lensing of the primordial perturbations, which will be analyzed in detail in a companion paper~\\cite{LENS}. It is also known that the RS effect scales as the third power of the comoving radius of the Void $L$ times the present value of the Hubble parameter $H_0$, $\\Delta T/T \\propto - (L H_0)^3$. For Void sizes compatible with the expectations from the usual structure formation scenarios, the RS effect happens to be suppressed with respect to the primordial temperature fluctuations ~\\cite{Fullana}. In order to produce a signal comparable to the measured temperature anisotropy of $|\\Delta T/T| \\sim 10^{-5}$, the Void should be of very large size, {\\it i.e.} $L \\sim (200-300) {\\rm Mpc}/h$: this would be at odds with the standard scenario of structure formation from Gaussian primordial fluctuations (at more than $10\\sigma$~\\cite{InoueSilk}). We may also mention that according to~\\cite{manyvoids,Szapudi} there are many localized regions in the sky (both underdense and overdense) of about $100 {\\rm Mpc}/h$, which would be responsible for the correlations between the CMB and the Large Scale Structures: as~\\cite{SarkarVoids} has recently stressed the existence of these regions is already at odds with the usual structure formation scenario. Even though we do not address in this paper the issue of the primordial origin of such large objects, we mention some possibilities. For example, one could consider certain models of inflation, such as the ones of the ``extended'' type~\\cite{extended,extendedDMN,extendedBN}, with the possibility of tunneling events and nucleation of bubbles, which would appear as Voids in the sky today. Alternatively, one could also imagine the presence of non-Gaussian features in the primordial fluctuations which would seed a large Void or even non-conventional structure formation histories in the late-time Universe which could enhance the probability of having large Voids. In the case of an explanation of the Voids via tunneling events, it is interesting to note that if a nucleation process has small probability (per unit volume and time) compared to the Hubble rate (at some time during inflation) then the number of anomalous Voids could well be very small, such as having just one or a few of them in the present observable Universe. In addition, since the LSS is a rather thin shell whose volume is much smaller than the total volume inside it, if there are only few Voids it is more likely that they are located along the line of sight, rather than at the LSS\\footnote{It is possible, also, to imagine that some primordial process could generate a coherent spherical region on the LSS surface, with small density contrast, on top of the primordial Gaussian spectrum, and which would produce a $10^{-5}$ fluctuation. However, we do not consider here such possibilities.}. This is an important point because a Void at the LSS would have a huge impact on the CMB and, as a consequence, its size would be strongly constrained: $L\\lesssim 4 {\\rm Mpc}/h$~\\cite{emptyVoids,extendedDMN}. On the contrary, Voids on the line of sight are not subject to the latter constraint. Note also that other authors have put forward alternative ideas, such as the idea that the Cold Spot could be due to a topological defect, in particular a cosmic texture~\\cite{texture}. As we are going to discuss, some of our considerations apply in a similar fashion if the origin of the Cold Spot is a texture, rather than a Void. The paper is organized as follows. In sect.~\\ref{profiles} we calculate the RS-induced temperature profile of the CMB photons passing through a large Void with the physical characteristics of the Cold Spot. The impact on the power spectrum is discussed in sect.~\\ref{powerspectrum}, while sect.~\\ref{bispectrum} deals with the impact on the bispectrum and on the overestimation of $f_{NL}$ that would be done by neglecting the presence of such a Void. Finally, in sect.~\\ref{multivoids} we extend some of the considerations to the case of several objects in the sky. In fact, it is conceivable that if there is one such structure, there may be other ones, maybe of smaller size and less visible -- see {\\it e.g}.~\\cite{manyvoids, Szapudi} for some other candidates in the CMB sky. Our conclusions are drawn in sect.~\\ref{concl}. ", "conclusions": "\\label{concl} Motivated by the so-called Cold Spot in the WMAP data, we have studied in this paper the impact on statistical CMB predictions, in particular the two and three point correlation functions, of the presence of an anomalously large Void along the line of sight. Indeed, the existence of such a Void could be at the origin of the Cold Spot, due to the Rees-Sciama effect (the Lensing effect is analyzed in the companion paper~\\cite{LENS}). First, we have computed the temperature profile using an LTB solution of the Einstein equations, matched to an FLRW metric. Then, we have computed its impact on statistical predictions for the CMB, {\\it assuming} this structure to be uncorrelated with the Primordial fluctuations. As suggested by~\\cite{texture}, we consider the angular size of such a Void to be about $10^\\circ-18^\\circ$ and its temperature at the centre to be characterized by $\\Delta T=-(190\\pm 80) \\mu K$. For the power spectrum the results are as follows. The RS effect is non-negligible: % we predict a bump of $5 \\% -25\\%$ to be added to the Primordial spectrum, localized at multipoles $5 \\leq \\ell \\leq 50$. This should lead to a variation in the $\\chi^2$ for the WMAP fits of order unity. Then we have studied the impact on the bispectrum coefficients. For the RS effect we have found that the Signal-to-Noise ratio is larger than unity at $\\ell \\gtrsim 40$ for most of the parameter space, and therefore this should already be visible in the available data. Through the bispectrum, we have studied also the impact of such a structure on the determination of the primordial non-gaussianity. The RS bispectrum signal is large but localized at low multipoles ($10 \\leq \\ell \\leq 50 $), so it has a small impact on high resolution experiments, which can go up to very large multipoles: the overestimation of $f_{NL}$ due to the RS effect turns out to be $\\Delta f_{NL}^{(RS)}\\simeq 1$ for WMAP and $\\Delta f_{NL}^{(RS)}\\simeq 0.1$ for Planck. Using the already existing WMAP1-year constraints~\\cite{WMAP1GAUSS} on $f_{NL}$ at low $\\ell$, one can exclude extreme values of the temperature contrast of the Void. For example, values for $\\Delta T/T$ larger than $8\\times 10^{-5}$ for a Cold Spot with diameter of $18^\\circ$ are likely to be excluded, via a full analysis. So, we conclude that the bispectrum is a valuable tool for constraining an anomalously large Void, through the RS effect. We have also considered the 50 Voids whose detection has been claimed in~\\cite{Szapudi}. These Voids have mean angular diameter of about $10^\\circ$ and average temperature at the centre characterized by $\\Delta T \\approx -11 \\mu K$. The effect on the two-point correlation function is about $0.5\\%$, hence much smaller than the one due to the large Void considered to explain the Cold Spot. The effect on the three-point correlation function has not been studied in detail but it is expected to lead to a Signal-to-Noise ratio smaller than one. Finally, we have also applied our considerations to the case in which the Cold Spot is explained by a cosmic texture~\\cite{texture} rather than a large Void: the effect on the power spectrum is similar but somewhat smaller; the three-point correlation function leads also to a smaller Signal-to-Noise ratio, which however turns out to be above unity for some part of the parameter space." }, "0808/0808.2435_arXiv.txt": { "abstract": "We have carried out a spatial-kinematic study of three proto-planetary nebulae, IRAS 16594$-$4656, Hen 3-401, and Rob 22. High-resolution \\mh images were obtained with NICMOS on the {\\it HST} and high-resolution spectra were obtained with the Phoenix spectrograph on Gemini-South. IRAS 16594$-$4656 shows a ``peanut-shaped'' bipolar structure with \\mh emission from the walls and from two pairs of more distant, point-symmetric faint blobs. The velocity structure shows the polar axis to be in the plane of the sky, contrary to the impression given by the more complex visual image and the visibility of the central star, with an ellipsoidal velocity structure. Hen 3-401 shows the \\mh emission coming from the walls of the very elongated, open-ended lobes seen in visible light, along with a possible small disk around the star. The bipolar lobes appear to be tilted 10$-$15$\\arcdeg$ with respect to the plane of the sky and their kinematics display a Hubble-like flow. In Rob 22, the \\mh appears in the form of an ``S''-shape, approximately tracing out the similar pattern seen in the visible. \\mh is especially seen at the ends of the lobes and at two opposite regions close to the unseen central star. The axis of the lobes is nearly in the plane of the sky. Expansion ages of the lobes are calculated to be $\\sim$1600 yr (IRAS 16594$-$4656), $\\sim$1100 yr (Hen 3-401), and $\\sim$640 yr (Rob 22), based upon approximate distances. ", "introduction": "\\label{intro} Proto-planetary nebulae (PPNs) are objects in transition between the asymptotic giant branch (AGB) and planetary nebula (PN) stages of stellar evolution. Studies over the past decade, particularly imaging studies with the {\\it Hubble Space Telescope} ({\\it HST}), have shown that this transition stage is key to understanding the shaping of PNs \\citep{balick02}. In the interacting stellar winds model \\citep{kwok82}, which was later generalized to include an equatorial density gradient \\citep{balick87}, it was assumed that a fast wind from the central star interacted with the detached, slowly-moving remnant AGB envelope to shape the nebula. {\\it HST} images of PPNs show that many and perhaps most of them display a basic bipolar structure \\citep{ueta00,su01,sahai07,siod08}. These are shapes that appear to be further developed in the PN stage. In addition, a point symmetry is often seen; \\citet{sahai98a} attribute this to collimated jets, which may be episodic and change their direction, and which they advocate as the main shaping agents of PNs. Studying \\mh in PPNs is one of the best ways to probe the presence of a fast, perhaps collimated, post-AGB wind and its effect in shaping the nebula. \\mh is the main constituent of the detached circumstellar envelope around a PPN. While a fast wind (V$>$100 km s$^{-1}$) has the energy to dissociate the \\mh on direct impact, it can also produce shocks in the medium that move at a slower speed and collisionally excite the \\mh. Recent surveys have shown the presence of shock-excited \\mh in bipolar PPNs \\citep{garher02, kelly05}. Detailed high-resolution \\mh studies have been published of approximately a half dozen PPNs \\citep{sahai98, kastner01, cox03, davis05, hri06, vds08}. In this paper we present the results of a detailed \\mh emission study of three bipolar PPNs, IRAS 16594$-$4656 (``Water Lily Nebula''), Hen 3$-$401 (IRAS 10178$-$5958), and Rob 22 (IRAS 10197$-$5750). All three have relatively hot central stars for PPNs, B and A spectral types, and all possess large infrared excesses due to circumstellar dust \\citep{partha89,garlar99a}. All three are or would be classified as DUPLEX (``DUst Prominent, Longitudinally-EXtended'') nebulae in the classification system of \\citet[see also \\citet{siod08}]{ueta00}. Rob 22 has a nebula that looks like a pair of butterfly wings. At high spatial resolution, the lobes show a filamentary structure in visible light. In addition to bipolar structure, some point symmetry is apparent in the lobes. A large halo extends out to 25$\\arcsec$ \\citep{sahai99a}. The nebula appears to be viewed essentially edge on, with a dark lane obscuring the star in visible light. In the more detailed morphological classification system of \\citet{sahai07}, this object is classified as Bcw,ml,ps(s),h(e): bipolar nebula with closed lobes, central obscuring waist, minor lobes present, point-symmetric shape, and enlongated halo. The spectral type is A2~Ie, based on light reflected off the lobes of the nebula \\citep{allen78}. The circumstellar envelope appears to have a dual chemistry, with an oxygen-rich region evidenced by OH maser emission \\citep[OH284.18-0.79;][]{allen80} and crystalline silicate emission \\citep{mol02} and a carbon-rich region evidenced by PAH emission \\citep{mol97}. The nebula of Hen 3-401 possesses long, narrow lobes. It seems to be viewed at a slightly larger angle with respect to the plane of the sky, with the central star appearing faintly between the lobes \\citep{sahai99b}. The morphological classification of this object is Bow*(0.6),sk: bipolar nebula with open lobes, central obscuring waist with star evident (at 0.6 $\\mu$m), with a skirt-like structure \\citep{sahai07}. The spectral type of the star is Be, and the spectrum shows strong Balmer emission lines and lots of permitted and forbidden low-excitation emission lines \\citep{allen78, garlar99b}. The circumstellar envelope appears to be carbon-rich, since it shows CO emission \\citep{loup90, buj91} and PAH emission \\citep{partha01} but not OH \\citep{silva93} emission or crystalline silicates \\citep{partha01}. In IRAS 16594$-$4656, the star is clearly seen and relatively bright compared to the nebula, giving the impression that the nebula is at a larger orientation angle. The nebula has the appearance of a basic bipolar structure with a pronounced point symmetry, consisting of three pairs of oppositely-directed, slightly curved, thin lobes. Also seen is a smaller elliptical structure oriented from northwest to southeast \\citep{hri99}. The morphological classification of this object is Mcw*(0.6),an,ps(m,an),h(a): multipolar nebula with closed lobes, central obscuring waist with star evident (at 0.6 $\\mu$m), ansae present, point symmetry with two or more pairs of diametrically opposed lobes and with ansae, and a halo with centrosymmetric arcs \\citep{sahai07}. Its spectrum shows Balmer emission lines \\citep{garlar99a} and has been classified as B7 based on the H$\\alpha$ line \\citep{vds00}. \\citet{rey02}, on the basis of high-resolution spectra, fit the spectrum with T$_{\\rm eff}$=14,000 K and log{\\it g}=2.1, slightly hotter than $\\beta$ Ori (B8~I); this indicates a spectral type of B5-7 Ie. He also observed many emission lines. The nebula appears to be carbon rich, showing PAH features and displaying the 21 $\\mu$m and 30 $\\mu$m features seen in carbon-rich evolved stars \\citep{garlar99a,volk02}. \\citet{vds08} recently published a detailed kinematic and morphologic study of this object using H$_2$ emission, and we will refer to their results in the present study. We begin with an examination of the high-resolution \\mh images (see \\citet{sahai00} for preliminary images of Rob 22 and Hen 3$-$401 and \\citet{hri04} for preliminary images of IRAS 16594$-$4656), then present new high-resolution, long-slit \\mh spectra, and then follow with a discussion of the kinematics. From these, we are able to determine properties of the collimated winds in these three PPNs and how they have shaped the surrounding nebulae. ", "conclusions": "The high-resolution {\\it HST}-NICMOS \\mh images of these three PPNs have allowed us to determine the spatial location of their \\mh emission and the spatially-resolved \\mh spectroscopic observations have allowed us to determine the kinematics of the \\mh emitting regions. In all three cases, the systemic and expansion velocities are similar to those determined from the molecular-line CO or OH measurements. We find the following results for these three PPNs. {\\it IRAS 16594$-$4656}: While the V and H-band images show a complex multi-lobe structure, the \\mh images shows a clear bipolar, peanut shape, although with variations in density (``holes'') and structure \\citep[see][]{vds08}. The \\mh emission originates along the walls of the lobes (sides and ends), with fainter emission from more distant (ejected?) clumps. The PV diagram shows the \\mh to arise in an expanding ellipsoidal velocity structure, which is in contrast to the bilobes of the density structure. The kinematics indicates that the bipolar lobes are nearly in the plane of the sky (i$\\approx$10$\\arcdeg$); this differs from the earlier interpretations of the lobes as being at some intermediate orientation, but it is consistent with recent mid-IR imaging and near-IR polarization studies. The lobes are estimated to have an age of $\\sim$1600 yr. {\\it Hen 3-401}: The \\mh emission originates from the sides of lobes, which have open ends. The lobes are tilted somewhat to the plane of the sky ($\\sim$10$-$15$\\arcdeg$), with the western lobe moving toward us, and they show an increasing velocity with radial distance (Hubble flow). An estimated age for the lobes is $\\sim$1100 yr. The open ends and unhindered outflow may be the consequence of its higher degree of collimation (greater linear momentum). {\\it Rob 22}: The \\mh emission originate primarily from ends of the S shaped nebula and from regions of the S shape near the obscured central star. The nebula is nearly in the plane of the sky, consistent with the absence of a visible star due to obscuration by a disk. An age of $\\sim$630 yr is estimated for the lobes. \\mh surveys of PPNs have shown that \\mh emission is commonly found in those with a bipolar morphology. As can be seen from this study, the combination of \\mh high-resolution images and spatially-resolved, high-resolution spectra provides valuable insight into the structure and shaping mechanisms for these bipolar nebulae." }, "0808/0808.2603_arXiv.txt": { "abstract": "We analyze archived {\\it Chandra} and {\\it XMM-Newton} \\mbox{X-ray} observations of 536 Sloan Digital Sky Survey (SDSS) Data Release 5 (DR5) quasars (QSOs) at $1.7 \\le z \\le 2.7$ in order to characterize the relative UV and \\mbox{X-ray} spectral properties of QSOs that do not have broad UV absorption lines (BALs). We constrain the fraction of \\mbox{X-ray} weak, non-BAL QSOs and find that such objects are rare; for example, sources underluminous by a factor of 10 comprise $\\la$2\\% of optically-selected SDSS QSOs. \\mbox{X-ray} luminosities vary with respect to UV emission by a factor of $\\la$2 over several years for most sources. UV continuum reddening and the presence of narrow-line absorbing systems are not strongly associated with \\mbox{X-ray} weakness in our sample. \\mbox{X-ray} brightness is significantly correlated with UV emission line properties, so that relatively \\mbox{X-ray} weak, non-BAL QSOs generally have weaker, blueshifted \\ion{C}{4}~$\\lambda$1549 emission and broader \\ion{C}{3}]~$\\lambda$1909 lines. The \\ion{C}{4} emission line strength depends on both UV and \\mbox{X-ray} luminosity, suggesting that the physical mechanism driving the global Baldwin effect is also associated with \\mbox{X-ray} emission. ", "introduction": "} It has been known for some time that the UV and \\mbox{X-ray} luminosities of quasars (QSOs) are correlated \\citep[e.g.,][and references therein]{at82}, and recent studies have carefully quantified this relation across $\\approx$5 orders of magnitude in UV luminosity \\citep[e.g.,][]{sbsvv05, ssbaklsv06, jbssscg07}. Such studies inform ongoing efforts to understand the structure and physics of QSO nuclear regions, providing quantitative constraints on models of physical associations between UV and \\mbox{X-ray} emission. Because UV photons are generally believed to be radiated from the QSO accretion disk while \\mbox{X-rays} are produced in the disk corona \\cite[e.g.,][and references therein]{rn03}, the \\mbox{UV/X-ray} luminosity relation is an indication of the balance between accretion disks and their coronae. For example, a large fraction of intrinsically X-ray weak sources would suggest that coronae may frequently be absent or disrupted in QSOs. In this study, we combine results from recent optical/UV and \\mbox{X-ray} observations of hundreds of QSOs in order to constrain the fraction of sources that are anomalously \\mbox{X-ray} weak, and also to test for additional physical effects that contribute to the scatter observed in the relation between UV and \\mbox{X-ray} luminosities. Radio-loud QSOs are well known to be relatively \\mbox{X-ray} bright \\citep[e.g.,][]{wtgz87, blvsbbwg00}, while QSOs with broad UV absorption lines (BALs) are \\mbox{X-ray} faint \\citep[e.g.,][]{gsahfbftm95, blw00}. We wish to understand the emission processes of QSOs without the added complexity of the strong X-ray absorption associated with UV BAL outflows or the enhanced X-ray emission that may be associated with a radio jet. In order to accomplish this, BAL and radio-loud QSOs must be carefully removed from samples. This process can be complicated in cases where available spectra do not extend to wavelengths where strong UV BAL absorption may occur. In this study, we compare the UV and \\mbox{X-ray} properties of optically-selected, radio-quiet Sloan Digital Sky Survey \\citep[SDSS; e.g.,][]{y+00} QSOs that are known not to host BAL outflows along the line of sight. Careful screening enables us to study the ways in which these ``ordinary'' QSOs deviate from the general relation between UV and \\mbox{X-ray} luminosities. We quantify the scatter about the best-fit relation between UV and \\mbox{X-ray} luminosities, and in particular constrain the fraction of optically-selected QSOs that are intrinsically weak \\mbox{X-ray} emitters. Secondly, we search for correlations between UV emission/absorption properties and {\\it relative} \\mbox{X-ray} brightness, where the term ``relative'' indicates the observed \\mbox{X-ray} brightness compared to that expected for an average QSO with the same UV luminosity. Beyond the UV/X-ray luminosity relation, our findings place additional constraints on physical models that relate QSO UV and \\mbox{X-ray} emission processes. The significant amount of scatter in the \\mbox{X-ray} brightness of individual sources compared to that of ``average'' QSOs with the same UV luminosity \\citep[e.g.,][]{ssbaklsv06} indicates that some unmodeled physical factors influence the relation between UV and \\mbox{X-ray} luminosities. Besides the cases of radio-loud and BAL QSOs mentioned above, possible additional causes of scatter include additional \\mbox{X-ray} absorption that is not associated with UV BALs, intrinsically weak (or strong) \\mbox{X-ray} emission, and time variability. Studies of \\mbox{X-ray} brightness in QSOs use different methods to quantify UV absorption and \\mbox{X-ray} weakness (complicating comparisons between different samples), and intensive UV and \\mbox{X-ray} spectroscopic campaigns have been performed for only a few QSOs. As a result, we do not have strong constraints on how frequently \\mbox{X-ray} absorption, intrinsically faint \\mbox{X-ray} emission, and time variability cause \\mbox{X-ray} weakness in QSO samples. In the remainder of this section, we briefly review published cases involving these physical scenarios. This discussion includes all the non-BAL QSOs listed in \\citet{blw00} and five of the six sources from \\citet{ybsv98} with $\\alpha_{OX} < -2$.\\footnote{The sixth source, QSO~$0316-346$, was identified from optical spectra by \\citet{mppb88}. UV spectra are required to search for BAL absorption, but no UV spectra of this source are present in the MAST or HEASARC archives.} We also discuss cases of unusual X-ray weakness in recent studies of individual sources. \\subsection{X-Ray Absorption\\label{xAbsCauseXWSec}} Even for QSOs that do not show UV BAL absorption, strong \\mbox{X-ray} absorption may be the simplest explanation for \\mbox{X-ray} weakness. Mrk~304 (PG~$2214+139$) is an \\mbox{X-ray} weak source that is often classified as a Seyfert~1 galaxy. It does not show UV BAL absorption, although \\citet{lb02} found some evidence for moderate \\ion{C}{4} and \\ion{N}{5} line absorption. In the \\mbox{X-rays}, it is strongly absorbed by a multizone, ionized absorber with a total column density $N_H \\sim 10^{23}$~cm$^{-2}$ \\citep{pjgsrs04, bpf04}. Another \\mbox{X-ray} weak source, PG~$1126-041$, shows evidence for an ionized \\mbox{X-ray} absorber. However, low-velocity, broad \\ion{C}{4} absorption has been found in some observations, indicating that PG~$1126-041$ should be classified as a (variable) BAL QSO \\citep{wbwyw99}. \\subsection{Intrinsic X-Ray Weakness\\label{intrinsicCauseXWSec}} \\mbox{X-ray} weakness cannot always be attributed to absorption. The nearby ($z = 0.192$), narrow-line type 1 QSO PHL~1811 shows no evidence of BALs in its UV spectrum. With an upper limit for a neutral \\mbox{X-ray} absorbing column of $N_H < 8.7 \\times 10^{20}$~cm$^{-2}$, the source is not strongly \\mbox{X-ray} absorbed \\citep{lhjgcp07}. It has been consistently \\mbox{X-ray} weak since it was first observed almost 20 years ago with {\\it ROSAT}, although the \\mbox{X-ray} flux has varied by a factor $\\approx$5. This variation (small compared to the degree of \\mbox{X-ray} weakness) suggested to \\citet{lhjgcp07} that the \\mbox{X-ray} emission was not scattered into view, as the medium responsible for the scattering would need to be implausibly small. PHL~1811 therefore appears to be an intrinsically weak \\mbox{X-ray} emitter. The optical/UV spectrum of PHL~1811 is also unusual \\citep{lhjc07}. It is very blue and shows no forbidden or semi-forbidden line emission. The \\ion{C}{4} $\\lambda 1549$ line emission is weak by a factor of $\\approx5$ compared to the composite spectrum of \\citet{fhfcwm91}. In this study, we will test whether intrinsically \\mbox{X-ray} weak objects like PHL~1811 (which would have been flagged for SDSS spectroscopy based on its optical colors alone) are common, and whether these unusual emission line characteristics are associated with \\mbox{X-ray} weakness. Another source, PG~$1011-040$, is relatively \\mbox{X-ray} weak by a factor of $\\sim$10 \\citep{blw00, gblemwi01}. There is some evidence for \\ion{C}{4} absorption in the UV spectrum, but no strong indication of a BAL \\citep{blw00, lb02}. It appears to be relatively unabsorbed in \\mbox{X-rays}, with an upper limit on the absorbing column density of $N_H \\le 5\\times 10^{21}$~cm$^{-2}$. In contrast to PHL~1811, multiple observations of PG~$1011-040$ have not found much \\mbox{X-ray} luminosity variation \\citep{gblemwi01}. \\subsection{UV and X-Ray Variability\\label{varCauseXWSec}} Source variability may be an important factor contributing to the scatter in the UV/\\mbox{X-ray} luminosity relation. Additional scatter is introduced by the fact that optical/UV and \\mbox{X-ray} observations are usually taken some time apart, and this time may reach up to years. The narrow-line Seyfert~1 galaxy Mrk~335 has recently been reported to have decreased in \\mbox{X-ray} brightness by a factor of $\\approx$30 \\citep{gkg07}. The authors suggested the variation may have been caused by the onset of heavy \\mbox{X-ray} absorption, which may also lead to the appearance of UV BALs. PG~$0844+349$ has been observed to vary in the \\mbox{X-rays} by up to 60\\% on short (20~ks) time scales, and by a factor of $\\sim$10 on longer (multi-year) time scales \\citep[e.g.,][]{gblemwi01, bgbf03}. The narrow-line Seyfert 1 galaxy WPVS~007 has dimmed in the \\mbox{X-rays} by a factor of 100 or more, while a \\ion{C}{4} BAL appears to be forming in the UV spectrum of this source \\citep{gslkon07}. Finally, we note that our understanding of the causes of \\mbox{X-ray} weakness is limited by observing time scales. Over longer time scales, \\mbox{X-ray} weak QSOs may show different properties (such as long-term emission or absorption variation), and may even cease to be \\mbox{X-ray} weak. We have also noted in \\S\\ref{introSec} cases where UV BALs may be transient, and recent study has demonstrated that BALs evolve on multi-year (rest-frame) time scales \\citep[e.g.,][]{gbsg08}. For these reasons, we are only able to place limits on the extent of \\mbox{UV/X-ray} variability (\\S\\ref{uVXVarConstraintSec}). Future near-simultaneous, multi-wavelength observations of QSOs will improve on these constraints. Throughout this work we use a cosmology in which $H_0 = 70$~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_M = 0.3,$ and $\\Omega_{\\Lambda} = 0.7$. ", "conclusions": "} We have analyzed the relationship between \\mbox{X-ray} luminosity and UV properties of SDSS DR5 QSOs at redshift $1.7 \\le z \\le 2.7$. At these redshifts, we are able to identify \\ion{C}{4} BAL QSOs in our sample. This has enabled us to place strong constraints on the fraction of intrinsically \\mbox{X-ray} weak QSOs, as well as to search for additional trends beyond the well-studied relationship between UV and \\mbox{X-ray} luminosities. In particular, we find that: \\begin{enumerate} \\item{Non-BAL, radio-quiet QSOs which are very (intrinsically) \\mbox{X-ray} weak are rare in optically-selected samples, with $\\la$2\\% of SDSS QSOs having $\\Delta\\alpha_{OX} < -0.4$ and $\\la$1\\% of SDSS/BQS QSOs having $\\Delta\\alpha_{OX} < -0.67$. Figure~\\ref{binomialProbOfltDAOXFig} provides upper limits on the fraction of sources that may be relatively \\mbox{X-ray} weaker than a given value of $\\Delta\\alpha_{OX}$.} \\item{The rms of the $\\Delta\\alpha_{OX}$ distribution for radio-quiet, non-BAL QSOs is about $0.1$, corresponding to a factor of $\\approx$2 spread in \\mbox{X-rays} relative to that estimated from the UV luminosity. This places an upper limit on typical \\mbox{X-ray} variability with respect to the UV continuum.} \\item{While the amplitude of and relation between UV and X-ray variation in QSOs is not well-understood, estimates based on recent studies indicate that most, if not all, of the observed spread of $\\Delta\\alpha_{OX}$ can be attributed to variability.} \\item{The distribution of $\\Delta\\alpha_{OX}$ may not be normally distributed (at 97.5\\% confidence) in our sample $B$. Perhaps physical processes such as \\mbox{X-ray} absorption are significant in some sources, broadening the wings of the distribution.} \\item{We find no strong evidence that reddening or the presence of narrow \\ion{Mg}{2} absorption is related to \\mbox{X-ray} weakness for non-BAL QSOs.} \\item{UV emission line properties such as \\ion{C}{4}~EW, \\ion{C}{3}]~FWHM, and perhaps \\ion{C}{4} wavelength are correlated to relative \\mbox{X-ray} brightness.} \\item{The \\ion{C}{4} emission EW depends on both UV and \\mbox{X-ray} luminosity. The physics that drives the global Baldwin effect is apparently associated with \\mbox{X-ray} emission as well as UV emission.} \\item{Even after correcting for secondary trends (such as weak \\ion{C}{4} emission, which is associated with relative \\mbox{X-ray} weakness), objects that are as intrinsically \\mbox{X-ray} faint as PHL~1811 are rare.} \\end{enumerate} We have quantitatively shown that luminous X-ray emission is essentially a universal property of optically-selected QSOs. Future studies can test whether this result holds true for QSOs selected in other wavebands, such as the radio and infrared. New optical surveys will extend the luminosity range of our sample, while UV (and infrared) spectroscopy will allow BAL QSOs to be identified (and $L_{2500~\\mathring{A}}$ to be measured) for sources in a wider redshift range. The high detection rate for relatively short exposures in our sample has also demonstrated that wide, shallow X-ray surveys at high angular resolution are an effective way to study the X-ray properties of bright, optically-selected QSOs." }, "0808/0808.2868_arXiv.txt": { "abstract": "We report the measurement of $\\nu$-$e$ elastic scattering from \\bor\\ solar neutrinos with 3\\,MeV energy threshold by the Borexino detector in Gran Sasso (Italy). The rate of solar neutrino-induced electron scattering events above this energy in Borexino is $0.217\\pm 0.038 (stat)\\pm 0.008 (syst)$~cpd/100\\,t, which corresponds to $\\Phi^{\\rm ES}_{\\rm ^8B}$ = {2.4 $\\pm$ 0.4$\\pm$ 0.1}$\\times$10$^6$~cm$^{-2}$ s$^{-1}$, in good agreement with measurements from SNO and SuperKamiokaNDE. Assuming the \\bor\\ neutrino flux predicted by the high metallicity Standard Solar Model, the average \\bor\\ \\nue\\ survival probability above 3 MeV is measured to be 0.29$\\pm$0.10. The survival probabilities for \\ber\\ and \\bor\\ neutrinos as measured by Borexino differ by 1.9 $\\sigma$. These results are consistent with the prediction of the MSW-LMA solution of a transition in the solar \\nue\\ survival probability \\Pee\\ between the low energy vacuum-driven and the high-energy matter-enhanced solar neutrino oscillation regimes. ", "introduction": "\\label{sec:intro} Solar \\bor-neutrino spectroscopy has been so far performed by the water \\che\\ detectors KamiokaNDE, SuperKamiokaNDE, and SNO~\\cite{Hir89,SKII08,SKI05,SNO07}. The first two experiments used elastic $\\nu$-$e$ scattering for the detection of neutrinos, whereas SNO also exploited nuclear reaction channels on deuterium with heavy water as target. These experiments provided robust spectral measurements with $\\sim$5\\,MeV threshold or higher for scattered electrons; a recent SNO analysis reached a 3.5\\,MeV threshold~\\cite{SNO09}. We report the first observation of solar \\bor-neutrinos with a liquid scintillator detector, performed by the Borexino experiment~\\cite{BXD08,BX09} via elastic $\\nu$-$e$ scattering. Borexino is the first experiment to succeed in suppressing all major backgrounds, above the 2.614\\,MeV $\\gamma$ from the decay of \\tal, to a rate below that of electron scatterings from solar neutrinos. This allows to reduce the energy threshold for scattered electrons by \\bor\\ solar neutrinos to 3\\,MeV, the lowest ever reported for the electron scattering channel. To facilitate a comparison to the results of SuperKamiokaNDE \\cite{SKI05} and SNO D$_2$O phase \\cite{SNO07}, we also report the measured \\bor\\ neutrino interaction rate with 5\\,MeV threshold. Since Borexino also detected low energy solar \\ber\\ neutrinos~\\cite{BX07,BX08}, this is the first experiment where both branches of the solar \\pp-cycle have been measured simultaneously in the same target. The large mixing angle solution (LMA) of the MSW effect~\\cite{MSW} predicts a transition in the $\\nu_e$ survival probability from the vacuum oscillation regime at low energies to the matter dominated regime at high energies. Results on solar \\ber\\ and \\bor\\ neutrinos from Borexino, combined with prediction on the absolute neutrino fluxes from the Standard Solar Model~\\cite{BS07,Pen08,Ser09}, confirm that our data are in agreement with the MSW-LMA prediction within 1$\\sigma$. ", "conclusions": "" }, "0808/0808.0576_arXiv.txt": { "abstract": "To date, fully cosmological hydrodynamic disk simulations to redshift zero have only been undertaken with particle-based codes, such as {\\tt GADGET}, {\\tt Gasoline}, or {\\tt GCD+}. In light of the (supposed) limitations of traditional implementations of smoothed particle hydrodynamics (SPH), or at the very least, their respective idiosyncrasies, it is important to explore complementary approaches to the SPH paradigm to galaxy formation. We present the first high-resolution cosmological disk simulations to redshift zero using an adaptive mesh refinement (AMR)-based hydrodynamical code, in this case, {\\tt RAMSES}. We analyse the temporal and spatial evolution of the simulated stellar disks' vertical heating, velocity ellipsoids, stellar populations, vertical and radial abundance gradients (gas and stars), assembly/infall histories, warps/lopsideness, disk edges/truncations (gas and stars), ISM physics implementations, and compare and contrast these properties with our sample of cosmological SPH disks, generated with {\\tt GCD+}. These preliminary results are the first in our long-term Galactic Archaeology Simulation program. ", "introduction": "The ability to form and evolve (correctly!) a disk galaxy with the aid of massively parallel computers and optimised algorithms remains an elusive challenge for astrophysicists. Ameliorating the non-physical effects associated with overcooling, overmerging, angular momentum loss, and the capture of accurate phenomenological prescriptions for the sub-grid physics governing galaxy evolution (star formation, feedback, etc.) has been achieved through rapid advancements in both hardware and software algorithms, but their complete elimination has yet to be realised. Fully self-consistent cosmological hydrodynamic simulations of Milky Way-like disk galaxies, with sufficient resolution ($\\simlt$500~pc) to decompose and analyse various galactic sub-components (eg. halo, bulge, and thin + thick disks) have only really appeared over the past $\\sim$5 years (Sommer-Larsen et~al. 2003; Abadi et~al. 2003; Governato et~al. 2004,2007; Robertson et~al. 2004; Bailin et~al. 2005; Okamoto et~al. 2005). A common thread linking these studies is the use of a particle-based approach to representing and solving the equations of hydrodynamics - usually through the use of a smoothed particle hydrodynamics (SPH) code, such as {\\tt GADGET}, {\\tt Gasoline}, or {\\tt GCD+}. Where there is no disputing the impact that SPH has had on the field, it is important to be aware of both the strengths {\\it and} weaknesses of any specific approach - as O'Shea et~al. (2005) and Agertz et~al. (2007) have shown, both subtle {\\it and} overt differences can be introduced when employing a particle-based, as opposed to a mesh-based (or grid-based) approach (and \\it vice versa\\rm), when simulating galaxy formation and evolution. To address these concerns, we have initiated a long-term Galactic Archaeology Simulation programme aimed at complementing the aforementioned particle-based studies (including our own) with a comprehensive suite of simulations generated with a grid-based N-body + hydrodynamical code employing adaptive mesh refinement (AMR) - our software tool of choice has been {\\tt RAMSES} (Teyssier 2002). \\it These simulations represent (to our knowledge) the first to be generated through to redshift zero, with a grid code, within a fully cosmological and hydrodynamic framework.\\rm\\footnote{The beautiful simulations of Ceverino \\& Klypin (2008), generated with the {\\tt ART} grid code were not (again, to our knowledge) run to redshift zero.} In this contribution, we provide a brief summary of the methodology adopted, and highlight several {\\it preliminary} results associated with our analyses of the simulations' disk kinematics, chemistry, disk edges / truncations, and assembly / infall histories. \\vspace{-4mm} ", "conclusions": "We have realised (to the best of our knowledge) the first fully self-consistent cosmological hydrodynamic disk simulations to $z$=0 with a mesh code; the resolution attained is 435~pc. Several preliminary results include: \\begin{itemize} \\item the saturated vertical disk heating seen in semi-cosmological SPH simulations has not yet been clearly replicated in our cosmological simulations; \\item the neutral gas disks show ``edges'' at comparable column densities to those observed; ionised gas disks extend beyond the neutral gas, again in agreement with those observed; \\item the stellar surface brightness profiles show ``breaks'' in the exponential profiles, with associated increases (reddening) in the age (colour) of the stellar populations beyond the break, in agreement with observation; little evidence is seen for an associated break in the stellar surface density profile, also as inferred from observations; stars of the same age beyond and interior to the break do not appear to have the same metallicity, which may prove problematic for radial migration scenarios; \\item gas accretion is not smooth, but does appear to be more-of-less ``inside-out''; \\item the disk-halo ``circulation flux'' is $\\sim$10-50$\\times$ that of the ``infalling flux'' (again, consistent with the broad numbers associated with the Milky Way). \\end{itemize} Beyond the analysis of the extant simulations, we have a number of planned enhancements, including full chemical evolution / tagging, a ten-fold increase in the number of simulations (to examine scaling relations, environmental dependencies, and assembly history variations), a range of ISM physics implementations (various polytropic equations of state, blast wave parametrisations), quantifying warp and lopsidedness statistics (Mapelli et~al. 2008), 2d IFU Lick index-style maps, dusty radiative transfer, high-velocity clouds, radial gas flows, and detailed SPH vs AMR comparisons with identical initial conditions. \\vspace{-4mm}" }, "0808/0808.2090_arXiv.txt": { "abstract": "We explore large-scale hydrodynamics of H \\Rmnum{2} regions for various self-similar shock flows of a polytropic gas cloud under self-gravity and with quasi-spherical symmetry. We formulate cloud dynamics by invoking specific entropy conservation along streamlines and obtain global self-similar ``champagne flows\" for a conventional polytropic gas with shocks as a subclass. Molecular cloud cores are ionized and heated to high temperatures after the onset of nuclear burning of a central protostar. We model subsequent evolutionary processes in several ways and construct possible self-similar shock flow solutions. We may neglect the mass and gravity of the central protostar. The ionization and heating of the surrounding medium drive outflows in the inner cloud core and a shock travels outwards, leading to the so-called ``champagne phase\" with an expanding outer cloud envelope. Complementarily, we also consider the expansion of a central cavity around the centre. As the inner cloud expands plausibly due to powerful stellar winds, a cavity (i.e., `void' or `bubble') can be created around the centre, and when the cavity becomes sufficiently large, one may neglect the gravity of the central protostar. We thus present self-similar shock solutions for ``champagne flows\" with an expanding central void. We compare our solutions with isothermal solutions and find that the generalization to the polytropic regime brings about significant differences of the gas dynamics, especially for cases of $n<1$, where $n$ is a key scaling index in the self-similar transformation. We also compare our global polytropic self-similar solutions with numerical simulations on the expansion of H \\Rmnum{2} regions. We further explore other possible dynamic evolutions of H \\Rmnum{2} regions after the initiation of nuclear burning of the central protostar, for example asymptotic inflows or contractions far from the cloud centre and the ongoing infall around a central protostar. In particular, it is possible to use the downstream free-fall solution with shocks to describe the dynamic evolution of H \\Rmnum{2} regions shortly after the nascence of the central protostar. We also give an analysis on the invariant form of self-similar polytropic flows by ignoring self-gravity. ", "introduction": " ", "conclusions": "" }, "0808/0808.1570_arXiv.txt": { "abstract": "Superhorizon perturbations induce large-scale temperature anisotropies in the cosmic microwave background (CMB) via the Grishchuk-Zel'dovich effect. We analyze the CMB temperature anisotropies generated by a single-mode adiabatic superhorizon perturbation. We show that an adiabatic superhorizon perturbation in a \\lcdm universe does not generate a CMB temperature dipole, and we derive constraints to the amplitude and wavelength of a superhorizon potential perturbation from measurements of the CMB quadrupole and octupole. We also consider constraints to a superhorizon fluctuation in the curvaton field, which was recently proposed as a source of the hemispherical power asymmetry in the CMB. ", "introduction": "The finite age of the Universe implies the existence of a cosmological particle horizon beyond which we cannot observe. Inhomogeneities with wavelengths longer than the horizon are not completely invisible, however. The generation of large-scale temperature fluctuations in the cosmic microwave background (CMB) by superhorizon perturbations is known as the Grishchuk-Zel'dovich effect \\cite{GZ78}. Through this effect, measurements of the low-multipole moments of the CMB \\cite{COBE, WMAP1} place constraints on the amplitudes and wavelengths of superhorizon perturbations. A well-known application of the Grishchuk-Zel'dovich effect uses CMB observations to place a lower bound on the size of the nearly homogeneous patch that contains the observable Universe. This bound was first derived for an Einstein-de Sitter universe \\cite{GZ78, Turner91}, and then for an open universe \\cite{KTF94, GLLW95}. Most recently, an analysis of the WMAP first-year data \\cite{WMAP1} found that our nearly homogeneous patch of the Universe extends to 3900 times the cosmological horizon \\cite{CDF03}. All of these analyses considered a statistically isotropic distribution of power in superhorizon perturbations and then asked how large the wavelength of order-unity perturbations needed to be in order to be consistent with the observed CMB anisotropies. In this paper, we analyze the CMB anisotropies induced by a single superhorizon adiabatic perturbation mode rather than an isotropic distribution of superhorizon inhomogeneities. A single-mode superhorizon perturbation to the gravitational potential would naively be expected to generate a dipolar CMB anisotropy with an amplitude comparable to the perturbation amplitude across the observable Universe. This is not the case in an Einstein-de Sitter universe, however, because the intrinsic dipole in the CMB produced by the perturbation is exactly cancelled by the Doppler dipole induced by our peculiar motion \\cite{GZ78, Turner91, BL94}. We show that the same cancellation occurs for an adiabatic superhorizon perturbation in a flat universe with a cosmological constant ($\\Lambda$), cold dark matter (CDM), and radiation. The strongest constraints to the amplitude and wavelength of a single superhorizon mode therefore arise from measurements of the CMB quadrupole and octupole. These constraints are less stringent than those derived for modes in a realization of a random-phase random field because it is possible to choose the phase of a single sinusoidal perturbation in such a way that there is no resulting quadrupole anisotropy. Single-mode superhorizon perturbations have received attention recently \\cite{GHHC05, Gordon07, DDR07, EKC08} because they introduce a special direction in our Universe and could be responsible for observed deviations from statistical isotropy in the CMB \\cite{TOH03, OTZH04, LM05, Eriksen04, HBG04, Eriksen07, GE08}. In particular, we investigated in a recent paper \\cite{EKC08} how a superhorizon perturbation during slow-roll inflation can generate an anomalous feature of the CMB: the fluctuation amplitude on large scales ($\\ell \\lsim 40$) is 10\\% larger on one side of the sky than on the other side \\cite{Eriksen04, HBG04, Eriksen07}. We first considered a perturbation to the inflaton field, but we found that the perturbation required to generate the observed power asymmetry induces large-scale anisotropies in the CMB that are too large to be consistent with measurements of the CMB octupole. We then considered a multi-field model of inflation in which a subdominant field, called the curvaton, is responsible for generating primordial perturbations \\cite{Mollerach90, LM97, LW02, MT01}. We found that a superhorizon perturbation in the curvaton field can generate the observed power asymmetry without inducing prohibitively large CMB anisotropies. We will use Ref.~\\cite{EKC08} as an example of how one may apply the CMB constraints to single-mode superhorizon perturbations derived here. We begin in Section \\ref{sec:basics} by reviewing the Grishchuk-Zel'dovich effect for adiabatic perturbations. In Section \\ref{sec:potential}, we derive the CMB anisotropy induced by a sinusoidal superhorizon perturbation in the gravitational potential, as would arise from a sinusoidal inflaton fluctuation. We also show in Section \\ref{sec:potential} that a superhorizon adiabatic perturbation does not generate a large dipolar anisotropy in a \\lcdm universe because the leading-order intrinsic dipole anisotropy is cancelled by the anisotropy induced by the Doppler effect. A sinusoidal curvaton fluctuation generates a potential perturbation that is not sinusoidal, and we derive the constraints to single-mode perturbations to the curvaton field in Section \\ref{sec:curvaton}. We summarize our results in Section \\ref{sec:conclusions}. Finally, an analytic demonstration of the dipole cancellation in a \\lcdm universe is presented in Appendix \\ref{app:cancel}, and the cancellation is shown to occur in flat universes containing a single fluid with an arbitrary constant equation of state in Appendix \\ref{app:cancel2}. ", "conclusions": "\\label{sec:conclusions} Superhorizon perturbations generate large-scale anisotropies in the CMB through the Grishchuk-Zel'dovich effect \\cite{GZ78}. In this paper, we have derived the constraints to single-mode adiabatic superhorizon perturbations that arise from measurements of the CMB quadrupole and octupole. These constraints differ from those previously derived for an isotropic distribution of superhorizon perturbations \\cite{GZ78, Turner91, KTF94, GLLW95, CDF03} because the CMB anisotropies generated by a single-mode perturbation depend on the perturbation's phase. We started by considering a sinusoidal superhorizon gravitational potential perturbation with wavenumber $k\\ll H_0$. Since the leading-order term in the potential perturbation is proportional to $(\\vec{k}\\cdot\\vec{x})$, it would be expected to generate a dipolar anisotropy of comparable amplitude in the CMB through the Sachs-Wolfe effect. However, the superhorizon perturbation also gives us a velocity with respect to the CMB, and the resulting Doppler dipole exactly cancels the leading-order intrinsic anisotropy generated by the SW and ISW effects, provided that the perturbation is adiabatic. This cancellation was known to occur in an Einstein-de Sitter universe \\cite{GZ78, Turner91, BL94}, but we found that it also applies to flat \\lcdm universes with and without radiation, as well as in more exotic flat cosmological models. Due to this cancellation of the CMB dipole, the leading-order constraints on adiabatic superhorizon fluctuations arise from measurements of the CMB quadrupole and octupole. If the potential perturbation is sinusoidal, as would be created by a sinusoidal fluctuation in the inflaton, then putting ourselves at the node of the sine wave maximizes the difference in potential across the Universe while also eliminating the induced quadrupole anisotropy in the CMB. In this case, the CMB octupole provides the strongest constraint on the amplitude of the superhorizon perturbation: $\\Delta \\Psi \\lsim 0.095$, where $\\Delta \\Psi$ is the variation of the potential $\\Psi$ across the surface of last scattering. A fluctuation in a field that contains only a small fraction of the energy density of the Universe generates a smaller potential perturbation and, consequently, smaller CMB anisotropies. We consider a multi-field model of inflation in which a subdominant curvaton field generates primordial perturbations \\cite{Mollerach90, LM97, LW02, MT01}. For a given superhorizon fluctuation in the curvaton field, the measured values of the CMB quadrupole and octupole place upper bounds on the fraction of the total energy density contained in the curvaton field prior to its decay. Since a sinusoidal perturbation in the curvaton field generates a potential perturbation that is not sinusoidal, there is no value for the phase of the curvaton fluctuation that eliminates the induced CMB quadrupole for any superhorizon curvaton fluctuation. However, once the amplitude of the curvaton perturbation is specified, it is possible to choose a phase for which the induced CMB quadrupole vanishes. In this case, measurements of the CMB octupole still place an upper bound on the curvaton energy density, but this bound is significantly weaker than the bound from the CMB quadrupole that applies to curvaton fluctuations with different phases. Superhorizon perturbations have generated interest recently because they are a simple way to introduce a preferred direction in the Universe and may generate the deviations from statistical isotropy that have been observed in the CMB. In particular, in Ref. \\cite{EKC08}, we showed that a superhorizon perturbation to an inflationary field can generate the hemispherical power asymmetry found in the WMAP data \\cite{Eriksen04, HBG04, Eriksen07}. In this paper, we have demonstrated how the CMB constrains such superhorizon perturbations: the octupole constraint on $\\Delta \\Psi$ is sufficient to rule out an inflaton perturbation as the source of the observed power asymmetry, but it is possible to generate the observed power asymmetry with a superhorizon curvaton perturbation. These constraints may also be applied to other scenarios that invoke superhorizon perturbations. For instance, order-unity superhorizon fluctuations in the mean value of the curvaton may be a generic feature of the curvaton model \\cite{LM06}." }, "0808/0808.0007_arXiv.txt": { "abstract": "Cluster galaxies moving through the intracluster medium (ICM) are expected to lose some of their interstellar medium (ISM) through ram pressure stripping and related ISM-ICM interactions. Using high-resolution cosmological simulations of a large galaxy cluster including star formation, we show that the ram pressure a galaxy experiences at a fixed distance from the cluster center can vary by well over an order of magnitude We find that this variation in ram pressure is due in almost equal parts to variation in the ICM density and in the relative velocity between the galaxy and the ICM. We also find that the ICM and galaxy velocities are weakly correlated for in-falling galaxies. ", "introduction": "X-ray observations of clusters have shown that substructure in the intracluster medium (ICM) is common (e.g. Mohr, Mathiesen, \\& Evrard 1999; Schuecker et al 2001). In a sample of 470 clusters, Schuecker et al. (2001) measure substructure in more than 50{\\%} of their sample. Detailed examinations of nearby clusters like Perseus and Virgo have discovered substructure and/or asymmetry in both the temperature and density profiles of these clusters (e.g. Bohringer et al. 1994; Shibata et al. 2001; Churazov et al. 2003; Dupke \\& Bregman 2001; Furusho et al 2001). Even Coma, considered a relaxed cluster, has ICM irregularities (White, Briel \\& Henry 1993). The importance of substructure on cluster mass measurements has been examined (Mohr, Mathiesen \\& Evrard 1999; Bohringer et al 2000), which in turn affects the use of cluster measurements as cosmological constraints (Jeltema et al. 2005; Nagai et al. 2007). However, the importance of substructure in the ICM is rarely considered when studying ram pressure stripped galaxies. Common assumptions are that the ICM is static, has a smooth density profile, and is only dense enough very near the center of a cluster to affect galaxies. Treu et al. (2003), in their evaluation of possible environmental evolutionary mechanisms in Cl 0024 + 16, assume that ram pressure is only effective to 0.6 virial radii. Solanes et al. (2001) find HI deficiency in galaxies out to two Abell radii, but only discuss the possibility that these galaxies are on highly radial orbits that have already carried them through the cluster center. Previous simulations studying galaxy evolution in clusters use a static, smooth ICM profile when studying the orbits of galaxies in clusters (e.g. Vollmer 2001; Roediger \\& Br{\\\"u}ggen 2007; J{\\'a}chym et al. 2007). These authors use different galaxy orbits in order to sample a variety of galaxy velocities at a fixed ICM density. Although the use of simple assumptions is widespread, there is at least one possible case in which ICM substructure had to be invoked to explain observations of the Virgo galaxy NGC 4522, a galaxy with a truncated gas disk (Kenney et al 2004; Vollmer et al. 2004; Vollmer et al. 2006). NGC 4522 is located at a projected distance of 1 Mpc from the center of the Virgo cluster, and assuming a static ICM with standard density values, the ram pressure is not strong enough to cause the observed truncation. Thus, the authors propose that the nearby ICM is either moving relative to the galaxy or overdense. In a recent paper studying the environmentally-driven evolution of galaxies in clusters using a detailed cosmological simulation (Tonnesen, Bryan \\& van Gorkom 2007), we examined the evolution of cool gas (i.e. ISM) in galaxies within and around the cluster, demonstrating that most gas loss from galaxies was due to ISM-ICM interactions (i.e. ram pressure and related processes), rather than galaxy-galaxy interactions or cluster tidal effects. We also found that ram pressure stripping occurs out to the virial radius of the cluster (measured using $r_{200}$). In this paper, we examine this result more closely and show that the ram pressure a galaxy experiences varies substantially, even at fixed distance from the cluster center. As we will see, this arises both from the density and velocity substructure of the ICM. First, we briefly introduce our code in \\S \\ref{sec:sim} and explain how we measure ram pressure in our simulation (\\S \\ref {sec:sample} and \\S \\ref{sec:rpsample}). We then present our results: a comparison of the standard deviations of ram pressure, ICM density, and velocity difference squared (\\S \\ref{sec:density}), followed by a more detailed look at the velocity of the ICM (\\S \\ref{sec:velocity}). ", "conclusions": "\\label{sec:discussion} In this paper we have presented a detailed examination of the intracluster medium with which a galaxy interacts as it falls into a simulated galaxy cluster. We find that substructure in the ICM is more important in varying ram pressure than is often assumed and used when modeling ram pressure stripping. Specifically, we highlight three main points: \\begin{enumerate} \\item In our simulated cluster we measure a range of ram pressure values for any given radius in the cluster. This ranges from an order of magnitude at 1 Mpc, to two orders of magnitude at the virial radius (1.8 Mpc), to even larger deviations further from the cD. Therefore, ram pressure can be effective at larger radii wherever there is an overdensity. \\item The scatter in ram pressure at different distances from the cD is due equally to the variation in the ICM density and the relative ICM-galaxy velocity ($v_{\\Delta}^2$) within the virial radius. This is true even when considering only higher ram pressure values. In fact, the normalized standard deviation in galaxy velocity is smaller than that of $v_{\\Delta}^2$. It is therefore not only the variety of orbital velocities that causes different values of ram pressure at fixed cluster radius, but also the density and velocity structure of the ICM. Further from the cD, ICM density variations dominate those of $v_{\\Delta}^2$. \\item The ICM velocity is correlated with galaxy velocity, resulting in a smaller $v_{\\Delta}$ than $v_{galaxy}$. This indicates that the ICM tends to move with in-falling galaxies, which then experience somewhat less ram-pressure than one would expect from a static ICM (although this is less true for high-velocity galaxies that are likely near the cluster center). \\end{enumerate} We emphasize that although we determine the ICM properties from a simulation, it is well-known in the X-ray cluster field that ICM substructure in density is common and in good agreement with simulations (Mohr, Mathiesen, \\& Evrard 1999; Jeltema et al. 2005; Nagai et al. 2007). Because ram pressure stripping is a fast process, even an overdensity with a relatively small extent can strip a galaxy that might otherwise retain its gas, or strip a galaxy more than predicted by its cluster position. Our results should galvanize the community currently studying galaxy evolution in clusters to look more closely at the intracluster medium." }, "0808/0808.2975_arXiv.txt": { "abstract": "The presence of atomic gas mixed with molecular species in a ``molecular'' cloud may significantly affect its chemistry, the excitation of some species, and can serve as probe of the cloud's evolution. Cold neutral atomic hydrogen (HI) in molecular clouds is revealed by its self absorption of background galactic HI 21-cm emission. The properties of this gas can be investigated quantitatively through observation of HI Narrow Self-Absorption (HINSA). In this paper, we present a new technique for measuring atomic gas physical parameters from HINSA observations that utilizes molecular tracers to guide the HINSA extraction. This technique offers a significant improvement in the precision with which HI column densities can be determined over previous methods, and it opens several new avenues of study of relevance to the field of star formation. ", "introduction": "Star formation occurs in molecular clouds which are thought to have evolved from diffuse atomic hydrogen (HI) regions to form dense, cold, well--shielded regions composed primarily of molecular Hydrogen (\\H2). Our quantitative understanding of cloud evolution and specifically the conversion process (from HI to \\H2 gas) has been hindered by our inability to measure confidently the HI abundance in evolving clouds. In this paper we present a new technique for measuring HI column densities in dark clouds offering a significant improvement over previous methods. In the interests of brevity this paper includes the results from only a few clouds on which this technique has been applied in order to demonstrate the technique. The results of a much larger survey of observational data and analysis are to follow in a subsequent publication. The ability to determine accurately the HI component of molecular clouds could have a variety of benefits. Measurement of HI/\\H2 ratios in clouds, used in conjunction with astrochemical models, allows us to determine the chemical ages of individual clouds or entire molecular complexes (e.g. Taurus, Perseus, etc.). This will greatly expand our understanding by constraining star formation models thus yielding insights into the collapse process and the interplay of magnetic fields, ambipolar diffusion, turblence, and various potential sources of cloud support. By studying age distributions in large scale regions we can learn about the processes that may trigger the collapse of large complexes. In contrast to previous methods, our technique allows us to determine the HI/\\H2 ratios for individual velocity components within a cloud thus yielding unique information about cloud kinematics. Further, the technique allows for the absolute measurement of quantities such as HI column density, in contrast to many previous studies which were limited to comparative measurements. HI is the dominant constituent of the diffuse ISM and its 21cm emission line is prevalent everywhere throughout the sky, especially near the galactic plane. Typical HI emission spectra are composed of numerous superimposed velocity components. The emission linewidths of the overall features are typically on the order of a few 10s of \\kms. They include velocity variations owing to galactic rotation as well as very significant peculiar velocities resulting from localized phenomena. It is rare to find a molecular region for which one can confidently claim that the HI emission observed along the line of sight to the cloud is associated with the cloud, and is not due to background or foreground sources. Owing to such velocity crowding it is difficult or impossible to disentangle HI emission originating from a particular cloud from the background galactic HI emission. The situation for HI absorption is different. HI within a galactic cloud of any type can absorb the continuum emission from distant (galactic or extragalactic) radio sources. Because the HI optical depth varies inversely with the cloud temperature, the absorption by galactic HI is stronger in cold, galactic HI clouds. The integrated optical depth of all the clouds along the line of sight through the galactic disk can exceed unity as demonstrated in \\cite{Kolpak}, and \\cite{Garwood}. These cold, interstellar HI clouds may also be identified through their absorption of warm background HI emission originating within the galaxy. Because the emission being absorbed by cold HI clouds in this case is galactic HI emission, the resultant spectral absorption features are refered to as HI self-absorption (HISA). There have been many surveys with HISA detections over the years including \\cite{garzoli}, \\cite{Heiles}, \\cite{Knapp}, \\cite{Heiles2}, \\cite{Wilson}, \\cite{McCutcheon}, \\cite{Myers}, \\cite{Bowers}, \\cite{Batrla}, \\cite{Shuter}, \\cite{vanderWerf}, \\cite{Feldt}, \\cite{Montgomery}, \\cite{Gibson}, and \\cite{Kavars03}, among others. It is important to emphasize that the term HISA refers to an observable spectral absorption feature rather than being a description of a specific physical process. Molecular spectral emission lines provide an independent view of that subset of interstellar clouds that are cold and composed primarily of molecular species. These ``molecular clouds'' are expected to maintain a residual abundance of atomic hydrogen if for no other reason than the cosmic ray disassociation of \\H2. This applies even when the chemical evolution of the cloud has reached equilibrium \\citep{Solomon1971, HINSA2}. With the residual HI co--existing throughout the cloud with molecular species, the observed spatial and velocity (kinematic) structure of the molecular cloud will be similar whether observed in molecular emission or HI self-absorption lines. Thus, we expect to observe HI self-absorption features along the lines of sight of molecular clouds which share the spatial distribution and kinematics (non--thermal line width) of molecular emission lines. Such localized association between molecular emission and HI self-absorption is observed for many nearby clouds as reported by \\cite{HINSA1} and \\cite{HINSA2}. The specific case in which the HI absorption features observed in the direction of a molecular cloud share the spatial and kinematic structure seen in the molecular lines is called HI Narrow Self-Absorption (HINSA) as defined in \\cite{HINSA1}. The term {\\it Narrow} arises from the typically small nonthermal linewidths (on the order of 0.1 \\kms) of HINSA features, very similar to the similarly small nonthermal linewidths of molecular tracers along the same lines of sight. HINSA can be considered to be a subset of HISA, but it is a subset derived from an understanding of a specific, observable physical phenomenon in molecular clouds. While both are simply acronyms for spectral absorption features, HISA can be caused by a variety of different conditions and processes, but HI Narrow Self-Absorption (HINSA) is that subset of HISA in which the atomic HI absorption correlates well with molecular emission of certain tracers (most notably $^{13}$CO) in sky position, central velocity, and nonthermal line width. Based on our current understanding of cold molecular clouds, the most satisfactory picture is that HINSA features are a result of HI gas located within these cold, dense, well--shielded regions. Some early examples of HINSA studies that predate the use of this term are those of \\cite{Wilson}, \\cite{vanderWerf}, and \\cite{Jackson}. The technique which we describe in this paper pertains only to the extraction of HI data from HINSA features. The general picture which emerges, as found in \\cite{HINSA1} and \\cite{HINSA2}, is that the HI gas located within cold, quiescent cores of dark clouds produces HINSA absorption features. The well--defined center velocities and narrow line widths allow us to separate the HI gas associated with individual clouds from the galactic background. However, the complexity of the background emission spectra that are frequently encountered makes extracting accurate data (especially in terms of obtaining the cold HI column density) from the absorption features difficult. Several methods (discussed in \\S \\ref{technique}) have been used previously, but all are recognized to introduce significant uncertainties in the results. We here present a new technique that aims to improve the situation by using the properties of molecular emission to characterize the region producing the HINSA features, and then employes the HINSA spectral features to derive HI column densities. In \\S \\ref{real} we present selected data and show the results of applying the new technique to them, and in \\S \\ref{previous} we contrast this with previous methods for analyzing HINSA data. In \\S \\ref{technique} and \\S \\ref{beamsizes} we describe the technique and the combination of molecular data with the HINSA spectra. In \\S \\ref{examples} and \\S \\ref{ambiguity} we verify the validity of our technique using simulated data and also examine its limitations. ", "conclusions": "\\label{summary} The significance of accurately measuring the HI content of dark clouds and star forming regions has long been recognized. HINSA features have been shown as a promising method to achieve this goal, but the difficulty in confidently disentangling HINSA from the galactic background emission has limited research efforts in this field. The technique described here builds upon previous work to provide new opportunities. By utilizing a second derivative representation in which HINSA becomes the dominant feature in the spectrum, and using information from associated molecular tracers to constrain our fits, we are able to obtain HI column densities with greater confidence than previously possible. This technique enables us to study individual velocity components within a cloud. Several uncertainties and sources of error still remain, most notably the errors inherent in temperature determination through $^{12}$CO, and the precise relations between the properties of cold HI and molecular tracers such as $^{13}$CO. As shown in \\S \\ref{examples} the purely statistical errors derived from our constructed data were only a few percent in simple cases which represent the majority of observed spectra. The nonlinearity error may be as large as 50 percent in the extreme cases. While the scope of this paper is limited to a demonstration of the new technique, the improved confidence in the results have the potential to yield significant scientific advances in the field of molecular cloud and star formation studies which are deferred to a subsequent publication. The two astronomical sources discussed briefly here for demonstration purposes are part of a much larger survey of over 30 dense cloud cores. In conjunction with numerical modeling these data provide us with the chemical ages of the clouds and individual components therein, thus providing a significant constraint on theoretical models of star formation. Large maps have been collected with the Arecibo L-Band Feed Array (ALFA) of the Taurus and Perseus star forming regions. These regions show plentiful HINSA features whose analysis may shed light on the dynamics of such complexes including the processes which may have triggered their formation." }, "0808/0808.0141_arXiv.txt": { "abstract": "We present new spectroscopic data for twenty six stars in the recently-discovered Canes Venatici~I (CVnI) dwarf spheroidal galaxy. We use these data to investigate the recent claim of the presence of two dynamically inconsistent stellar populations in this system~\\citep{Ibata2006}. We do not find evidence for kinematically distinct populations in our sample and we are able to obtain a mass estimate for CVnI that is consistent with all available data, including previously published data. We discuss possible differences between our sample and the earlier data set and study the general detectability of sub-populations in small kinematic samples. We conclude that in the absence of supporting observational evidence (for example, metallicity gradients), sub-populations in small kinematic samples (typically fewer than 100 stars) should be treated with extreme caution, as their detection depends on multiple parameters and rarely produces a signal at the 3$\\sigma$ confidence level. It is therefore essential to determine explicitly the statistical significance of any suggested sub-population. ", "introduction": "\\label{sec:intro} It is now widely accepted that the dwarf spheroidal (dSph) satellite galaxies of the Milky Way and Andromeda are the most dark matter dominated stellar systems known in the Universe~\\citep[e.g.][]{Mateo1998}. Over the past two decades, a significant amount of observational work has focussed on quantifying both the amount of dark matter in these systems, and its spatial distribution~\\citep[e.g.][]{Gilmore2007,Walker2007}. Although recent numerical simulations have shown that many of the dSphs may not be immune to tidal disturbance by the Milky Way~\\citep[e.g.][]{Munoz2008,Lokas2008}, their observed properties still require the presence of massive dark matter haloes which protect them against complete tidal disruption. The dSphs thus provide us with nearby laboratories in which to test dark matter theories. Given that dSphs occupy the low luminosity end of the galaxy luminosity function, their star formation histories provide useful insights into the star formation process. Analyses of spatial variations in colour-magnitude diagram morphology provided early evidence of population gradients in a number of dSphs~\\citep[e.g.][]{Harbeck2001}. More recently, evidence of metallicity gradients has been found using spectroscopic estimates of [Fe/H]~\\citep[e.g.][]{Tolstoy2004,Koch2006,Battaglia2006}. In at least one case, that of the Sculptor dSph, the metal-rich and metal-poor populations have significantly different spatial distributions and kinematics~\\citep{Tolstoy2004,Battaglia2008}. Although little evidence of similar features has been found in other dSphs~\\citep[e.g.][]{Koch2006, Koch2007a, Koch2007b}, the presence of dynamically distinct stellar populations within dSphs, as well as the complex interplay between the dynamical, spatial and chemical properties of their stars, is of great interest as it has implications for star formation and galaxy evolution. It is, however, important to note that although the hierarchical build-up of structure in the standard $\\Lambda$-Cold Dark Matter ($\\rm{\\Lambda\\,CDM}$) paradigm implies that satellite galaxies contribute significantly to the stellar haloes of their hosts, detailed abundance studies of stars in the more luminous dSphs have demonstrated that their properties are significantly different from those of the Milky Way halo~\\citep[e.g.][]{Shetrone2001,Helmi2006}. Among the significant differences between the halo and the dSphs, the more important chemical differences are in the alpha-elements~\\citep{Unavane1996,Venn2004}. The observed gradients in the heavy element distributions are reproduced by the models of supernova feedback in dSphs developed by~\\cite{Marcolini2008}. Thus, it appears that the primordial dwarf satellites, which were disrupted to form the Milky Way halo, had stellar populations distinct from those seen in the present-day dSphs~\\citep{Robertson2005, Font2006}. Given their high estimated mass-to-light ratios, the observed dSphs are usually identified with the large population of sub-haloes which are observed to surround Milky Way-sized haloes in cosmological simulations assuming a standard $\\rm{\\Lambda\\,CDM}$ universe. However, it was noted early on that the number of dSphs around the Milky Way was much lower than the expected number of satellite dark matter haloes~\\citep[e.g.][]{Moore1999}. A number of possible explanations for the apparent lack of Milky Way satellites have been presented in the literature~\\citep[e.g.][]{Stoehr2002, Diemand2005, Moore2006, Strigari2007, Simon2007, Bovill2008}. All these models are based on the reasonable postulate that out of the full population of substructures around the Milky Way, the observed dSphs are merely the particular subset which (for reasons of mass, orbit, formation epoch, re-ionisation, etc.) were able to capture gas, form stars and survive any subsequent tidal interactions with the Milky Way. In addition, the ratio between the predicted and observed numbers of dwarf galaxies has decreased significantly in the past few years due to the discovery of nine new Milky Way dSph satellites~\\citep{Willman2005,Zucker2006a,Zucker2006b,Belokurov2006,Belokurov2007,Walsh2007} in the data from the Sloan Digital Sky Survey~\\citep[SDSS;][]{York2000}. Since the SDSS covers only about one fifth of the sky, it is thus likely that the total number of satellites surrounding the Milky Way may be at least a factor of five larger than previously thought, although the extrapolation from the SDSS survey to the whole sky requires careful analysis~\\citep[see e.g.][]{Tollerud2008} . In order to compare the properties of the newly discovered satellites with those of sub-haloes in cosmological simulations, as well as to confirm their nature as true satellite galaxies of the Milky Way, as opposed to star clusters or disrupted remnants, spectroscopic observations of their member stars are essential in order to estimate dynamical masses from the observed stellar kinematics. The extremely low luminosities of these objects ~\\citep[in some cases as low as $10^3$L$_\\odot$: ][]{Martin2008b}, present significant observational challenges as the kinematic data sets are small, making it difficult to obtain statistically significant results. The Canes Venatici~I (CVnI) dSph is the brightest of the newly discovered population of very faint SDSS dSphs~\\citep{Zucker2006a}. ~\\cite{Ibata2006} presented spectra for a sample of CVnI member stars obtained using the DEIMOS spectrograph mounted on the Keck telescope. They identified two kinematically distinct stellar populations in this data set: an extended metal-poor population with high velocity dispersion and a centrally-concentrated metal-rich population with a dispersion of almost zero. Their analysis of the mass of CVnI suggested that the two populations might not be in equilibrium as the mass profiles obtained based on the individual populations were inconsistent with each other. However, a subsequent study of CVnI by ~\\cite{Simon2007}, using a larger sample of Keck spectra, did not reproduce this bimodality. An important outstanding question is whether the ultra-faint dSphs represent the low-luminosity tail of the dSph population, or are instead the brightest members of a population of hitherto unknown faint stellar systems, distinct from both dSphs and star clusters. The presence of multiple, distinct kinematic populations in a low-luminosity dSph would set it apart from the majority of low-luminosity star clusters. In addition, the presence of a spread in the stellar abundances would suggest an association with the brighter dwarf galaxies and would also be interesting in terms of its implications for star formation. It is thus important to determine whether the sub-population identified by ~\\cite{Ibata2006} in CVnI is real. One goal of our study was to shed some light on this issue by using spectra obtained with a different spectrograph to those in the previous two studies of CVnI. In addition, we wanted to investigate the extent to which sub-populations can be reliably detected in the very small kinematic data sets which are observable for the ultra-faint dSphs. In addition to their potential importance for probing the star formation histories of dSphs, kinematic substructures can be used to test another key feature of the hierarchical structure formation paradigm. The fact that dark matter clustering occurs on all scales means that the dSph satellites of the Milky Way are likely to be in the process of accreting their own population of smaller satellites. Although these substructures may not have been able to form their own stars, they may be able to acquire stars from their host dSph. They would then be detectable as localised populations with mean velocity and/or velocity dispersion distinct from that of the dSph. Populations with these properties have, in fact, been detected in the Ursa Minor and Sextans dSphs~\\citep{Kleyna2003,Walker2006}. Once a dSph halo begins to fall into the Milky Way, it will cease to accrete new satellites as any nearby substructures will rapidly be removed by the tidal field of the Milky Way and the high relative velocities in the Milky Way halo will preclude the capture of new satellites. Due to the short internal dynamical timescales in dSphs (typically a few hundred Myr), any remaining internal substructures will subsequently be destroyed on timescales of at most a few Gyrs if dSph haloes are cusped, although they can survive much longer if their haloes are cored~\\citep{Kleyna2003}. In the standard cusped-halo picture, only those satellites which have been interacting with the Milky Way for less than a few internal dynamical times, either because they are currently passing the Milky Way for the first time as may be the case for the Leo~I dSph~\\citep[]{Mateo2008} or the Magellanic Clouds ~\\citep[]{Kallivayalil2006, Besla2007,Piatek2008} or because their crossing times are larger~\\citep[e.g. the Magellanic Clouds:][]{vanderMarel2002}, would be expected to exhibit localised kinematic substructure. If localised substructures were found to be common in dSphs, this could be difficult to reconcile with a picture in which dSphs occupy cusped haloes. Given that the level of substructure above a given mass fraction is a function of halo mass \\citep{Gao2004}, the expected numbers of sub-haloes per dSph requires further investigation by means of cosmological simulations. However, the importance of comparing the level of substructure in dSphs with the results of numerical simulations adds further motivation to our goal of establishing the level of confidence with which sub-populations can be detected in small data sets. The outline of the paper is as follows. In Section~\\ref{sec:cvn}, we present a new kinematic data set for stars in CVnI, based on spectra obtained with the Gemini telescope, and calculate a mass estimate for the galaxy from these data. In Section~\\ref{sec:pop}, we look for kinematic sub-populations in our data, and compare our findings with those of ~\\cite{Ibata2006}. Section~\\ref{sec:det} discusses the general detectability of sub-populations in small kinematic data sets. Finally, in Section~\\ref{sec:conc} we draw some general conclusions and suggest possible differences between the two data sets for CVnI that we have compared. ", "conclusions": "\\label{sec:conc} In this paper, we have presented a new data set of velocities and metallicities for the Canes Venatici I (CVnI) dSph based on spectra taken with the GMOS-North spectrograph. A maximum likelihood fit to the velocity distribution yields a mean velocity of $v= 25.8\\pm0.3$\\,km\\,s$^{-1}$ and a dispersion of $\\sigma =7.9^{+1.3}_{-1.1}$\\,km\\,s$^{-1}$. Assuming a constant, isotropic velocity dispersion and a Plummer profile for the mass distribution, we find a mass of $4.4^{+1.6}_{-1.1}\\times10^7 M_\\odot$ in the volume where our tracer stars are located. Although this value is larger than the value $2.7\\pm 0.4\\times 10^7 M_\\odot$ calculated by ~\\cite{Simon2007},this is most likely due to the assumptions made for our models and the distribution of our particular subsets of stars. One of the original aims of our study was to investigate the claimed multiple stellar populations in CVnI. As we discussed above, the two previous studies by ~\\cite{Ibata2006} and ~\\cite{Simon2007} did not agree on the existence of a cold sub-population in CVnI. The two populations found in the former study were puzzling as they led to two different mass estimates. The authors suggested that this might indicate that the system had recently accreted a younger population and was not yet in equilibrium. In this paper we looked for evidence of multiple populations in our data under the assumption that each population was Gaussian. Based on this analysis, we concluded that there was no reason to suspect the presence of a second population in our data. We also applied our analysis to the \\cite{Ibata2006} data where we found evidence of a statistically significant sub-population with a dispersion of $\\sigma =0.6$\\,km\\,s$^{-1}$ (compared to $\\sigma =13.6$\\,km\\,s$^{-1}$ for the main population). Our analysis suggests that there is a qualitative difference between our data and those of \\cite{Ibata2006}. Although further data would be necessary to resolve this issue, we note that the spatial distributions of these two data sets are different, which could potentially account for the differences in the detected populations. However, our central field is centred close to the blue/young star population which \\cite{Martin2008a} find in their photometry from the Large Binocular Telescope, and which they identify with the cold population of \\cite{Ibata2006}. The exact fraction of stars in each population found by \\cite{Martin2008a} is currently unclear, however, and so it is possible that we have not picked up any stars associated with the cold population. We have also carried out a study of the detectability of sub-populations in small kinematic data sets. Under the assumption of Gaussian populations, we studied the effects of four parameters. We obtained confidence limits for the detection of sub-populations in samples with different numbers of stars, different population ratios and velocity dispersions. We found that reasonable errors on the observed velocities do not affect the detectability of the sub-populations. For a given sample size, our ability to detect two populations increased as the ratio of their dispersions $\\sigma_{1}/\\sigma_{2}$ increased. However, even for large $\\sigma_{1}/\\sigma_{2}$ and equal population size, a sample of 30 stars yielded a $3\\sigma$ detection in only $\\sim35$ per cent of cases. As expected, for larger sample sizes, this detection rate was significantly higher. We also showed that a cold population needs to constitute a larger fraction of the total sample than is required to detect a hot sub-population. This suggests that the robust detection of the sub-populations associated with any surviving sub-haloes within a dSph would require samples of many hundreds of velocities. In this case, localised substructures could be detected by windowing the data, provided that a window whose spatial size coincided with plausible sub-halo scales would contain a sample of at least 100 stars. As such data sets are now becoming available for many of the larger dSphs, this test may soon be feasible. We note that the claim of multiple global populations in Sculptor~\\citep[]{Tolstoy2004} was based on a large data set and is therefore still robust. Finally, we note that all our significance tests were based on the assumption of Gaussian populations, which was the case for all our Monte Carlo samples. However, for real data, the true distributions will not be known, and are not necessarily well-approximated by Gaussians. It is therefore difficult in a real case to assign a robust statistical significance to a particular detection of a sub-population. As we have shown, for small data sets, many Monte Carlo realisations do not yield significant detections of the sub-populations. In the absence of a robust estimate of the confidence level of a particular detection, or additional, independent evidence of the presence of multiple populations, we conclude that one should exercise great caution in decomposing data sets of fewer than $100$ stars into multiple populations." }, "0808/0808.0188_arXiv.txt": { "abstract": "{In recent years, increasing evidence for chemical complexity and multiple stellar populations in massive globular clusters (GCs) has emerged, including extreme horizontal branches (EHBs) and UV excess.} % {Our goal is to improve our understanding of UV excess in compact stellar systems, covering the regime of both ultra-compact dwarf galaxies (UCDs) and massive GCs.} {We use deep archival GALEX data of the central Fornax cluster to measure NUV and FUV magnitudes of UCDs and massive GCs. } {We obtain NUV photometry for a sample of 35 compact objects that cover a range $-13.5-$1) GC candidates in M31 with significant FUV flux which are thought to be analogs of two peculiar Galactic GCs, NGC 6388 and NGC 6441 (Yoon et al. 2008). Sohn et al. (2006) and Kaviraj et al. (2007) analysed the UV properties of massive globular clusters associated with M87 in the Virgo cluster, and found that many of them show a UV-excess with respect to canonical stellar population models. These findings support the idea that EHBs may be a common feature to the most massive compact stellar systems. In this Research Note we focus on the UV properties of compact stellar systems in the Fornax cluster. In contrast to the studies of Sohn et al. and Kaviraj et al. on Virgo GCs, we extend our analysis to compact stellar systems beyond the mass range of GCs ($M\\lesssim 3 \\times 10^6$ M$_{\\sun}$), including the so-called ultra-compact dwarf galaxies (UCDs, Drinkwater et al. 2003), which cover the mass range up to $\\sim10^8$M$_{\\sun}$, having $M_V \\lesssim -11$ mag. We analyse how the UV properties of UCDs compare to those of both massive and normal GCs, in order to improve our knowledge of EHB occurence in compact stellar systems. Throughout this study we adopt (m-M)=31.4 mag (Freedman et al. 2001) as distance modulus to the Fornax cluster. \\begin{figure*}[] \\begin{center} \\epsfig{figure=f1a.eps,width=8.6cm} \\epsfig{figure=f1b.eps,width=8.6cm} \\caption{{\\bf Left:} Map of the central Fornax cluster. The large circles indicate the FoV of the two archival GALEX pointings used for this study. The large dotted circle corresponds to the Deep Imaging Survey (DIS), the large dashed circle corresponds to the Near Galaxies Survey (NGS). The small dots indicate our sample of spectroscopically confirmed compact objects down to $V\\simeq 22$ mag ($M_V=-9.4$ mag). The filled circles indicate the sources with visually verified GALEX matches in the NUV, the asterisks are those sources with matches in the FUV (matching radius 3$''$). {\\bf Right:} V,I colour-magnitude diagram of the same sources as in the left panel. The optical data is from Mieske et al. (2004 \\& 2006, 2007), and Dirsch et al. (2003). Magenta dots indicate GCs in M31 with UV colours available (Fig.~\\ref{cmds}; Rey et al. 2007), red circles indicate GCs in M87 with FUV colours available (Fig.~\\ref{cmds}; Sohn et al. 2006). Green squares indicate sources with X-ray counterparts from the Chandra study of Scharf et al. (2005). Note that only one of the GALEX UV detections has a detected X-ray counterpart.} \\label{map} \\end{center} \\end{figure*} ", "conclusions": "A UV excess in an old stellar population is likely due to EHB stars. As pointed out in Sect.~\\ref{introduction}, an EHB may be linked to helium-enriched stars (e.g. Ventura et al. 2001, D'Antona et al. 2002). The strong UV excess of the seven massive Fornax GCs beyond the He-enhanced isochrones, especially the NUV excess of the three most extreme GCs with (NUV-V)$<$2.4 mag (see Fig.~\\ref{colcol}), suggests that in these objects, EHB formation is also driven by other processes. In this context, a plausible explanation may be enhanced mass loss of evolved stars, triggered by high stellar densities (Decressin et al. 2007; Huang \\& Gies 2006) and/or large binary fractions. Excess radiation at short wavelengths can in principle also arise from accretion onto a black hole (King et al. 1993), which can be traced by low-mass X-ray binaries (Jord\\'{a}n et al. 2004). We have cross-checked the positions of all GALEX UV detections with X-ray source detections in the Chandra Fornax Survey data (Scharf et al. 2005 and private communication), the deepest available wide-field X-ray survey of Fornax (50ks integration with ACIS). The sensitivity of these images is a few $ 10^{38}$ erg/sec, allowing to detect the most luminous LMXBs (Jord\\'{a}n et al. 2004). In Fig.~\\ref{map} { (right panel)} we indicate the (V-I) optical colours of those compact objects with X-ray matches. At a given magnitude, the X-ray matches { in GCs} are biased towards red optical colours (see also Jord\\'{a}n et al. 2004), while GALEX UV detections are biased towards blue optical colours. This suggests that generally, the UV- and X-ray-emission of the compact stellar systems are not caused by the same physical processes. However, there is one GALEX UV detection with an X-ray counterpart (Fig.~\\ref{map} and~\\ref{colcol}), which happens to be one of the three GCs with largest UV excess. We can therefore not exclude that the UV excess in some of the GCs is linked to accretion processes. We finally note that comparing the probability of UV excess between UCDs and GCs allows to test whether EHBs are more likely associated with present-day deep potential wells (i.e. UCDs) or high stellar densities (i.e. GCs; Dabringhausen et al. 2008, Mieske et al. 2008). One would expect deep potential wells to favour self-enrichment (e.g. Ventura et al. 2001, D'Antona et al. 2002), and high stellar densities to favour mass-loss scenarios (Decressin et al. 2007; Huang \\& Gies 2006). Such a comparison may therefore help to constrain the efficiency of EHB formation channels, provided that the present-day density and mass of the systems investigated have not experienced significant changes during the past, which could have been the case due to core collapse (Noyola \\& Gebhardt 2006, de Marchi et al. 2007) or tidal stripping (e.g. Lee et al. 2007). To properly perform this comparison, deeper UV imaging data will be required that allow detection of UV intermediate-bright to faint GCs down to $M_V\\simeq-$10 mag ($V\\simeq 21.5$ mag at the Fornax distance). In this respect, the outcomes of the HST observations in Cycle 15 (GO10901, PI O'Connell) of GCs belonging to NGC 1399 are highly anticipated." }, "0808/0808.3374_arXiv.txt": { "abstract": "At a distance of about 130~pc, the Corona Australis molecular cloud complex is one of the nearest regions with ongoing and/or recent star formation. It is a region with highly variable extinction of up to $A_{V} \\sim 45$ mag, containing, at its core, the \\textsl{Coronet} protostar cluster. There are now 55 known optically detected members, starting at late B spectral types. At the opposite end of the mass spectrum, there are two confirmed brown dwarf members and seven more candidate brown dwarfs. The CrA region has been most widely surveyed at infrared wavelengths, in X-rays, and in the millimeter continuum, while follow-up observations from centimeter radio to X-rays have focused on the \\textsl{Coronet} cluster. ", "introduction": "{\\em Corona Australis (southern crown)}, abbreviated {\\em CrA}, is one of the 48 ancient constellations listed by Ptolemy, who called it {\\em Stephanos Notios} (in Greek). The latin translation was {\\em corona notia, austrina, australis, meridiana,} or {\\em meridionalis}, from which \\citet{all36} picked {\\em corona australis}; see also \\citet{bot80} for a discussion of the name. For an overview, see Fig.~\\ref{lokekuntan}. \\subsection{The Discovery} \\label{sec_disc} Given the convention for naming variable stars, R CrA was the first variable star noticed in the CrA constellation. Marth from Malta and Schmidt from Athens (quoted in \\citealp{rey16}) were the first to notice the variability of the nebula NGC 6729 in CrA; Schmidt also discovered the variable stars R, S, and T CrA with the nebula attached to R CrA. The variability of stars and nebula were confirmed by Innes at Cape Observatory (quoted in \\citealp{rey16}) and \\citet{kno16}; the latter gave $\\Delta m$ = 1.4 mag for the variability of R CrA in images taken from 21 Aug 1911 to 23 May 1915 at Helwan Observatory (Egypt). \\citet{rey16} also noticed that the behaviour of R CrA (variability and environment) is similar to that of the star T Tau. \\citet{luy27} found the three bright variables UZ CrA, VW CrA, and VX CrA. Then, \\citet{vge32,vge33} found more than 100 variable stars in or near CrA, based on plates taken at the Union Observatory in Johannesburg, mostly by himself and Hertzsprung, a few of them young stars. \\citet{her60} estimated the age of the two early-type emission-line stars R CrA (Ae) and T CrA (F0e) as their main sequence contraction time ($10^{5}$ to $10^{7}$ years) and concluded that they are young, and that therefore also the associated nebula and the other stars should be young. \\citet{kna73} found infrared (IR) excess in some of the variable stars in CrA, indicative of circumstellar material, and derived $\\sim 1$ Myr as the age of the group. The Corona Australis molecular cloud complex is today known as one of the nearest regions with ongoing and/or recent intermediate- and low-mass star formation. The dark cloud near R CrA is the densest cloud core in the region with extinction of up to $A_{V} \\sim 45$ mag \\citep{dau87,wil92}. This cloud is also called {\\em condensation A} \\citep{ros78}, {\\em FS~445 -- FS~447} \\citep{fes84}, DoH {\\em 2213} \\citep[][see Fig.~\\ref{dobashifig}]{dob05}, and sometimes also just {\\em CrA} or {\\em R CrA cloud}; we use {\\em CrA} for {\\em Corona Australis} as the name for the star-forming region. \\begin{figure*} \\includegraphics[width=\\linewidth]{cra_fig1.eps} \\caption{Stars and dust in CrA. Credit: Loke Kun Tan (StarryScapes).} \\label{lokekuntan} \\end{figure*} \\begin{figure*} \\centering \\includegraphics*[width=0.95\\linewidth,bb=20 54 576 546]{cra_fig2.eps} \\caption{The CrA cloud in an extinction map from \\citet{dob05}. Note the location relative to the Galactic plane at $b\\approx-18^\\circ$.} \\label{dobashifig} \\end{figure*} \\begin{figure*} \\includegraphics*[width=\\linewidth,bb=19 23 575 266]{cra_fig3.eps} \\caption{Extinction map of CrA from $B$-band star counts (from \\citealp{cam99}).} \\label{cambresy_fig6} \\end{figure*} \\subsection{The Galactic Environment} The CrA dark cloud is located $\\sim 18^{\\circ}$ below the Galactic plane (see Fig.~\\ref{dobashifig}). According to \\citet[see also Fig. 1.10 in \\citealp{poe97}]{ola82}, the CrA cloud is not part of the Gould Belt nor of the Lindblad ring \\citep{poe97}, but is located within the massive HI shell called {\\em Loop I} \\citep{har93}. The 3D space motion of the CrA T~Tauri stars would be consistent with the dark cloud being formed originally by a high-velocity cloud impact onto the Galactic plane, which triggered the star formation in CrA \\citep{neu00}. When a high-velocity cloud impacts onto the gas and dust in the Galactic plane, this high-velocity cloud may get distroyed, but then forms a new (lower velocity) cloud. Alternatively, the star formation in CrA may have been caused by the expansion of the UpperCenLupus (UCL) superbubble: \\citet{mam01} found that the CrA complex is radially moving away from UCL with 7~km~s$^{-1}$ so that it was located near its center $\\sim 14$~Myrs ago. \\citet{har93} resolved cloud {\\em A} into five condensations, the star R~CrA being located in {\\em A2}. \\citet{anv96} and \\citet{cam99} mapped the cloud using optical star counts (see Fig.~\\ref{cambresy_fig6}). \\citet{juv08} study the relationship between scattered near-infrared light and extinction in a quiescent filament of the CrA cloud and compare the result to extinction measurements. Several IR surveys revealed a large population of embedded IR sources \\citep{tas84,wil86,wil89,wil92,wil97} some of which are IR Class~I objects. Based on their $^{13}$CO map, \\citet{tac02} estimated the star formation efficiency in CrA to be $\\sim 4~\\% $. \\subsection{The Distance} The distance towards the CrA star forming region was estimated by \\citet{gap36} to be $150 \\pm 50$~pc and later by \\citet{mar81} to be $\\sim 129$~pc (assuming a ratio of total to selective absorption of $R=4.5$, see \\citealp{vrb81}). The \\textsl{Hipparcos} satellite attempted to measure the parallax of the star R CrA and found $122 \\pm 68$ mas, i.e. no reliable solution. \\citet{knh98} gave $\\sim 170$ pc as distance. Three likely CrA members have reliable \\textsl{Hipparcos} parallax measurements and thus distances; these are HR~7170 ($d$=$85\\pm26$~pc), SAO~210888 ($d$=$193\\pm39$~pc), and HD~176386 ($d$=$139\\pm22$~pc). R~CrA and HR~7169 do not have reliable parallax measurements. The weighted mean of these three distances is $128\\pm26$~pc. \\citet{cas98} determined the distance towards the eclipsing double-lined spectroscopic binary TY CrA to be $129 \\pm 11$~pc from their orbit solution. The orbit solution of TY~CrA provides the currently best estimate for the distance to the star-forming region, and we therefore suggest to use a distance of 130~pc for CrA, see also \\citet{dez99}. ", "conclusions": "" }, "0808/0808.3835_arXiv.txt": { "abstract": "% NGC\\,2264 is a young Galactic cluster and the dominant component of the Mon OB1 association lying approximately 760 pc distant within the local spiral arm. The cluster is hierarchically structured, with subclusters of suspected members spread across several parsecs. Associated with the cluster is an extensive molecular cloud complex spanning more than two degrees on the sky. The combined mass of the remaining molecular cloud cores upon which the cluster is superposed is estimated to be at least $\\sim$3.7$\\times$10$^{4}$ M$_{\\odot}$. Star formation is ongoing within the region as evidenced by the presence of numerous embedded clusters of protostars, molecular outflows, and Herbig-Haro objects. The stellar population of NGC\\,2264 is dominated by the O7 V multiple star, S~Mon, and several dozen B-type zero-age main sequence stars. X-ray imaging surveys, H$\\alpha$ emission surveys, and photometric variability studies have identified more than 600 intermediate and low-mass members distributed throughout the molecular cloud complex, but concentrated within two densely populated areas between S~Mon and the Cone Nebula. Estimates for the total stellar population of the cluster range as high as 1000 members and limited deep photometric surveys have identified $\\sim$230 substellar mass candidates. The median age of NGC\\,2264 is estimated to be $\\sim$3 Myr by fitting various pre-main sequence isochrones to the low-mass stellar population, but an apparent age dispersion of at least $\\sim$5 Myr can be inferred from the broadened sequence of suspected members. Infrared and millimeter observations of the cluster have identified two prominent sites of star formation activity centered near NGC\\,2264 IRS1, a deeply embedded early-type (B2--B5) star, and IRS2, a star forming core and associated protostellar cluster. NGC\\,2264 and its associated molecular clouds have been extensively examined at all wavelengths, from the centimeter regime to X-rays. Given its relative proximity, well-defined stellar population, and low foreground extinction, the cluster will remain a prime candidate for star formation studies throughout the foreseeable future. ", "introduction": "Few star forming regions surpass the resplendent beauty of NGC\\,2264, the richly populated Galactic cluster in the Mon OB1 association lying approximately 760 pc distant in the local spiral arm. Other than the Orion Nebula Cluster, no other star forming region within one kpc possesses such a broad mass spectrum and well-defined pre-main sequence population within a relatively confined region on the sky. Estimates for the total stellar population of the cluster range up to $\\sim$1000 members, with most low-mass, pre-main sequence stars having been identified from H$\\alpha$ emission surveys, X-ray observations by {\\it ROSAT}, {\\it Chandra}, and {\\it XMM-Newton}, or by photometric variability programs that have found several hundred periodic and irregular variables. The cluster of stars is seen in projection against an extensive molecular cloud complex spanning more than two degrees north and west of the cluster center. The faint glow of Balmer line emission induced by the ionizing flux of the cluster OB stellar population contrasts starkly with the background dark molecular cloud from which the cluster has emerged. The dominant stellar member of NGC\\,2264 is the O7 V star, S~Monocerotis (S~Mon), a massive multiple star lying in the northern half of the cluster. Approximately 40\\arcmin\\ south of S~Mon is the prominent Cone Nebula, a triangular projection of molecular gas illuminated by S~Mon and the early B-type cluster members. NGC\\,2264 is exceptionally well-studied at all wavelengths: in the millimeter by Crutcher et al. (1978), Margulis \\& Lada (1986), Oliver et al. (1996), and Peretto et al. (2006); in the near infrared (NIR) by Allen (1972), Pich\\'{e} (1992, 1993), Lada et al. (1993), Rebull et al. (2002), and Young et al. (2006); in the optical by Walker (1956), Rydgren (1977), Mendoza \\& G\\'omez (1980), Adams et al. (1983), Sagar \\& Joshi (1983), Sung et al. (1997), Flaccomio et al. (1999), Rebull et al. (2002), Sung et al. (2004), Lamm et al. (2004), and Dahm \\& Simon (2005); and in X-rays by Flaccomio et al. (2000), Ramirez et al. (2004), Rebull et al. (2006), Flaccomio et al. (2006), and Dahm et al. (2007). NGC\\,2264 was discovered by Friedrich Wilhelm Herschel in 1784 and listed as H VIII.5 in his catalog of nebulae and stellar clusters. The nebulosity associated with NGC\\,2264 was also observed by Herschel nearly two years later and assigned the designation: H V.27. The Roman numerals in Herschel's catalog are object identifiers, with `V' referring to very large nebulae and `VIII' to coarsely scattered clusters of stars. One of the first appearances of the cluster in professional astronomical journals is Wolf's (1924) reproduction of a photographic plate of the cluster and a list of 20 suspected variables. Modern investigations of the cluster begin with Herbig (1954) who used the slitless grating spectrograph on the Crossley reflector at Lick Observatory to identify 84 H$\\alpha$ emission stars, predominantly T Tauri stars (TTS), in the cluster region. Herbig (1954) postulated that these stars represented a young stellar population emerging from the dark nebula. Walker's (1956) seminal photometric and spectroscopic study of NGC\\,2264 discovered that a normal main sequence exists from approximately O7 to A0, but that lower mass stars consistently fall above the main sequence. This observation was in agreement with predictions of early models of gravitational collapse by Salpeter and by Henyey et al. (1955). Walker (1954, 1956) proposed that these stars represent an extremely young population of cluster members, still undergoing gravitational contraction. Walker (1956) further noted that the TTSs within the cluster fall above the main sequence and that they too may be undergoing gravitational collapse. Walker (1956) concluded that the study of TTSs would be ``of great importance for our understanding of these early stages of stellar evolution.'' \\begin{figure}[!tbh] \\begin{center} \\includegraphics[angle=90,width=\\linewidth, draft=False]{fg1.eps} \\caption[fg1.eps]{A 12.5$^{\\circ}$ square false-color IRAS image (100, 60 \\& 25~$\\mu$m) of the Mon OB1 and Mon R1 associations. NGC\\,2264 lies at the center of the image with several nearby IRAS sources identified, including the reflection nebulae NGC\\,2245 and NGC\\,2247, and NGC\\,2261 (Hubble's Variable Nebula). South of NGC\\,2264 is the Rosette Nebula and its embedded cluster NGC\\,2244, lying 1.7 kpc distant in the Perseus arm. \\label{f1}} \\end{center} \\end{figure} The molecular cloud complex associated with NGC\\,2264 was found by Crutcher et al. (1978) to consist of several cloud cores, the most massive of which lies roughly between S~Mon and the Cone Nebula. Throughout the entire cluster region, Oliver et al. (1996) identified 20 molecular clouds ranging in mass from $\\sim10^{2}$ to $10^{4}$ M$_{\\odot}$. With NGC\\,2264 these molecular clouds comprise what is generally regarded as the Mon OB 1 association. Active star formation is ongoing within NGC\\,2264 as evidenced by the presence of numerous embedded protostars and clusters of stars, as well as molecular outflows and Herbig-Haro objects (Adams et al. 1979; Fukui 1989; Hodapp 1994; Walsh et al. 1992; Reipurth et al. 2004a; Young et al. 2006). Two prominent sites of star formation activity within the cluster are IRS1 (also known as Allen's source), located several arcminutes north of the tip of the Cone Nebula, and IRS2, which lies approximately one-third of the distance from the Cone Nebula to S~Mon. New star formation activity is also suspected within the northern extension of the molecular cloud based upon the presence of several embedded IRAS sources and giant Herbig-Haro flows (Reipurth et al. 2004a,c). From 60 and 100~$\\mu$m IRAS images of NGC\\,2264, Schwartz (1987) found that the cluster lies on the eastern edge of a ring-like dust structure, 3$^{\\circ}$ in diameter. Shown in Figure 1 is a 12\\fdg5$\\times$12\\fdg5 false-color IRAS image (100, 60, and 25~$\\mu$m) centered near NGC\\,2264. The reflection nebulae NGC\\,2245 and NGC\\,2247, members of the Mon R1 association, are on the western boundary of this ring (see the chapter by Carpenter \\& Hodapp). Other components of the Mon R1 association include the reflection nebulae IC\\,446 and IC\\,2169, LkH$\\alpha$215, as well as several early type (B3--B7) stars. It is generally believed that the Mon R1 and Mon OB 1 associations are at similar distances and are likely related. The Rosette Nebula, NGC\\,2237-9, and its embedded young cluster NGC\\,2244 lie 5$^{\\circ}$ southwest of NGC\\,2264, 1.7 kpc distant in the outer Perseus arm (see the chapter by Rom\\'an-Z\\'u\\~niga \\& Lada). Several arcs of dust and CO emission have been identified in the region, which are believed to be supernovae remnants or windblown shells. Many of these features are apparent in Figure 2, a wide-field H$\\alpha$ image of NGC\\,2264, NGC\\,2244, and the intervening region obtained by T. Hallas and reproduced here with his permission. It is possible that star formation in the Mon OB1 and R1 associations was triggered by nearby energetic events, but it is difficult to assess the radial distance of the ringlike structures evident in Figure 2, which may lie within the Perseus arm or the interarm region. Shown in Figure 3 is a narrow-band composite image of NGC\\,2264 obtained by T.A. Rector and B.A. Wolpa using the 0.9 meter telescope at Kitt Peak. S~Mon dominates the northern half of the cluster, which lies embedded within the extensive molecular cloud complex. \\begin{figure} \\centering \\includegraphics[angle=0,width=\\linewidth, draft=False]{fg2.eps} \\caption[fg2.eps]{An extraordinary widefield, narrow-band H$\\alpha$ image of NGC\\,2264 (upper center), the Rosette Nebula and NGC\\,2244 (lower right), and the numerous windblown shells and supernova remnants possibly associated with the Mon OB1 or Mon OB2 associations. The Cone Nebula is readily visible just above and left of image center as is S~Mon. Also apparent in the image is the dark molecular cloud complex lying to the west of NGC\\,2264. This image is a composite of 16 20-minute integrations obtained by T. Hallas using a 165 mm lens and an Astrodon H$\\alpha$ filter. \\label{f2}} \\end{figure} To summarize all work completed over the last half-century in NGC\\,2264 would be an overwhelming task and require significantly more pages than alloted for this review chapter. The literature database for NGC\\,2264 and its members has now grown to over 400 refereed journal articles, conference proceedings, or abstracts. Here we attempt to highlight large surveys of the cluster at all wavelengths as well as bring attention to more focused studies of the cluster that have broadly impacted our understanding of star formation. The chapter begins with a review of basic cluster properties including distance, reddening, age, and inferred age dispersion. It then examines the OB stellar population of the cluster, the intermediate and low-mass stars, and finally the substellar mass regime. Different wavelength regions are examined from the centimeter, millimeter, and submillimeter to the far-, mid-, and near infrared, the optical, and the X-ray regimes. We then review many photometric variability studies of the cluster that have identified several hundred candidate members. Finally, we consider future observations of the cluster and what additional science remains to be reaped from NGC\\,2264. The cluster has remained in the spotlight of star formation studies for more than 50 years, beginning with the H$\\alpha$ survey of Herbig (1954). Its relative proximity, low foreground extinction, large main sequence and pre-main sequence populations, the lack of intense nebular emission, and the tremendous available archive of observations of the cluster at all wavelengths guarantee its place with the Orion Nebula Cluster and the Taurus-Auriga molecular clouds as the most accessible and observed Galactic star forming region. \\begin{figure}[!tbh] \\vspace{8mm} \\begin{center} \\includegraphics[angle=0,width=\\linewidth, draft=False]{fg3.eps} \\caption[fg3.eps]{Narrow-band, three-color image of NGC\\,2264 obtained by T.A. Rector and B.A. Wolpa (NOAO/AURA/NSF) using the 0.9 meter telescope at Kitt Peak National Observatory. The filters used for this composite image are: O III (light blue), H$\\alpha$ (red-orange), and [S II] (blue-violet). The field of view is approximately 0.75$^{\\circ}$$\\times$1$^{\\circ}$. S~Mon lies just above the image center and is believed to be the ionizing source of the bright rimmed Cone Nebula. \\label{f2}} \\end{center} \\vspace{-8mm} \\end{figure} ", "conclusions": "" }, "0808/0808.3718_arXiv.txt": { "abstract": "We present a system of X-ray photometry for the \\chandra satellite. X-ray photometry can be a powerful tool to obtain flux estimates, hardness ratios, and colors unbiased by assumptions about spectral shape and independent of temporal and spatial changes in instrument characteristics. The system we have developed relies on our knowledge of effective area and the energy-to-channel conversion to construct filters similar to photometric filters in the optical bandpass. We show that the filters are well behaved functions of energy and that this X-ray photometric system is able to reconstruct fluxes to within about 20\\%, without color corrections, for non-pathological spectra. Even in the worst cases it is better than 50\\%. Our method also treats errors in a consistent manner, both statistical as well as systematic. ", "introduction": "\\label{sec:intro} Photometry is one of the most widely used, relatively simple, tools used in describing and categorizing astronomical objects. Standardization by \\citet{johnson:53} and subsequent additions in the optical and infrared (see e.g. \\citet{bessel:05}) have allowed comparisons between measurements by different telescopes and instruments without bias. Important applications of optical photometry include stellar classifications (see e.g. \\citet{johnson:53}), galaxy redshifts \\citep{puschell:81}, and the discovery of the most distant quasars in the SDSS \\citep{fan:99}. In particular photometry is important for sources too faint to extract detailed spectra, i.e. most sources, given the almost universal increase of source numbers to faint fluxes. Although optical astronomy is a much older discipline than X-ray astronomy, optical photometry was established only about 20 years before the beginning of X-ray astronomy in the 1960s. Given the enormous success of optical photometry it seems obvious to use it as a precedent and try to duplicate its success in other wavebands beyond UV and infrared. The X-ray band can be defined as reaching from about 0.1 keV to a few hundred keV, spanning almost 4 decades of frequency -- although most work concentrates on the 0.2--20 keV band -- compared with 2 octaves in the optical. Moreover, an important difference between optical and X-ray is the typically low number of photons in X-ray astronomy. The X-ray range is photon starved such that sources with a few hundred counts are considered bright in X-rays. This limitation increases the importance of broad-band photometry in the X-ray band. While X-ray astronomy has used relative, mission-specific photometry for most of its existence, there is, as yet, no standard X-ray photometric system. The usefulness of an X-ray photometric system is evident already from the use of these somewhat idiosyncratic energy bands. The bands used in the past have been chosen for specific purposes, e.g. to use color--color diagrams to diagnose X-ray binary spectral states \\citep{white:84,hasinger:89} where, in the latter, the energy bands are different for each source. Even so, the resulting color--color diagrams have immensely increased our knowledge of X-ray binary spectral/accretion states \\citep[e.g.][]{prestwich:03,gierlinski:06}. Thus a standard X-ray photometric system is highly desirable in X-ray astronomy in order to cross-compare observations of the hundreds of thousands of sources being cataloged by XMM \\citep{watson:07}, \\chandra \\citep{fabbiano:07}, and other missions. Even within a given mission different types of CCDs (XMM) or changes in operating temperature, gain or contamination (\\chandra) mean that simple count rates cannot be used. However, there have been complicating factors in establishing photometric energy bands beyond individual observations. One cause of this lack of standardization is that the energy ranges covered by different X-ray satellites and instruments differ widely. For example, RXTE/PCA, GINGA and EXOSAT bands have practically no overlap with ROSAT bands; and ASCA, XMM and \\chandra bands are somewhere in the middle. Fig. \\ref{fig:ebands} shows a selection of energy bands used by different authors for different X-ray satellites\\footnote{For more information on X-ray satellites, see the HEASARC web page http://heasarc.gsfc.nasa.gov/docs/observatories.html. A list of energy bands and references is given in the on-line version of this paper.}. Most X-ray missions with focusing optics cover the energy range from $\\sim$0.1 keV to 10 keV. Another reason for the lack of a standard photometric system is that in X-rays there are no bright constant point sources in the way that stars can be used for calibration like in the optical. But the most fundamental cause for the lack of an X-ray photometric system has been the limited spectral resolving power (R$=$E/$\\Delta E\\sim$1) of proportional counters, which were used in X-ray astronomy from the earliest days through to ROSAT and RXTE. A resolution of R$\\sim$1 allows no clean separation of energy bands, and different spectra with similar flux will give widely different flux estimates in any chosen band. In optical terms, the ``color correction'' is very large. However, with the introduction of X-ray CCDs in ASCA \\citep{burke:93} this limitation has largely gone away. X-ray CCDs have R$>$10, so comparable to the R$\\sim$6 of broad band optical photometry. It seems thus timely to consider the introduction of an X-ray photometric system. Therefore we have investigated how good a photometric system can be created for \\chandra ACIS observations and, by extension, for all other X-ray CCDs. We report the encouraging results in this paper. \\bfig[h] \\resizebox{\\hsize}{!}{\\includegraphics{f1.eps}} \\caption{Energy bands used in various publications for X-ray satellites. The different energy bands for different satellites are due to energy coverage of the instruments. The colors represent the width of the soft/medium/hard bands used in the corresponding papers. A table of energy bands and references are given in the on-line version of the paper.\\label{fig:ebands}} \\efig ", "conclusions": "We have presented a system of X-ray photometry for the \\chandra satellite. The system we have developed relies on a knowledge of effective area and the energy-to-channel conversion to construct X-ray filters, but is unbiased by assumptions about the spectral shape of a source. We have shown that the filters are comparable to filters in the optical and infrared, and that our photometric system in X-rays is able to estimate fluxes to within about 20\\%. Even in the worst cases it is better than 50\\%. We have incorporated methods to estimate systematic errors and consistently propagate statistical as well as systematic errors. Due to the construction method employed our filter system is very flexible and can be adapted readily to other CCD X-ray detectors, in particular to XMM-Newton EPIC. The code to compute fluxes is available at a web page \\footnote{http://hea-www.cfa.harvard.edu/~jcm/xray/index.html}. A table with a selection of correction factors for ACIS-S3 is given in Appendix A in the electronic edition and available at the same URL. The table contains only every fourth tile because of the generally slowly varying correction factor with chip location. In the future we will explore potential improvements to X-ray photometry by means of: \\begin{itemize} \\item Making color corrections using band ratios. A preliminary investigation suggests that extreme spectra (e.g. highly absorbed or high photon indexes) will gain significantly in the accuracy of flux estimates and even normal spectra will have a reduced error range. \\item Optimizing the choice of bands. The properties of our current method show that there is a correlation between the accuracy of flux estimates and the filter shape. The more ``boxy'' a filter is the better the flux estimate will be. This suggests a limit for the width of a filter at which point deviations from boxiness result in an accuracy of the flux estimate below a certain value. \\item Comparing results for \\chandra ACIS and XMM-Newton EPIC data. \\chandra and XMM have similar instrumental setups (ARF and RMF) and overlapping science capabilities. And given the large numbers of sources observed by these two X-ray missions it is very important to be able to compare results for the two. \\end{itemize}" }, "0808/0808.2414_arXiv.txt": { "abstract": "Broad absorption line quasars (commonly termed BALQSOs) contain the most dramatic examples of AGN-driven winds. The high absorbing columns in these winds, $\\sim$~$10^{24}$~cm$^{-2}$, ensure that BALQSOs are generally X-ray faint. This high X-ray absorption means that almost all BALQSOs have been discovered through optical surveys, and so what little we know about their X-ray properties is derived from very bright optically-selected sources. A small number of X-ray selected BALQSOs (XBALQSOs) have, however, recently been found in deep X-ray survey fields. In this paper we investigate the X-ray and rest-frame UV properties of five XBALQSOs for which we have obtained XMM-Newton EPIC X-ray spectra and deep optical imaging and spectroscopy. We find that, although the XBALQSOs have an $\\alpha_{ox}$ steeper by $\\sim$~0.5 than normal QSOs, their median $\\alpha_{ox}$ is nevertheless flatter by 0.30 than that of a comparable sample of optically selected BALQSOs (OBALQSOs). We rule out the possibility that the higher X-ray to optical flux ratio is due to intrinsic optical extinction. We find that the amount of X-ray and UV absorption due to the wind in XBALQSOs is similar, or perhaps greater than, the corresponding wind absorption in OBALQSOs, so the flatter $\\alpha_{ox}$ cannot be a result of weaker wind absorption. We conclude that these XBALQSOs have intrinsically higher X-ray to optical flux ratios than the OBALQSO sample with which we compare them. ", "introduction": "In seeking to understand any role that AGN winds might play in the coevolution of supermassive black holes and their host galaxies, we need to study the physical properties of those winds at z~$\\sim$~2 where star formation and black hole growth were at their height. It is probable that the majority of mass and energy carried in AGN winds is transported by the X-ray absorbing part of the outflow \\citep[e.g.][]{blustin2007}, as the X-ray absorber has the highest column density, at least in nearby AGN, but it is usually difficult to obtain satisfactory X-ray spectra from high-redshift AGN. The most well-known population of distant AGN showing evidence of ionised winds is the Broad Absorption Line quasars (BALQSOs; z~$\\sim$~0.1$-$6), whose winds have outflow speeds of up to 60000 km~s$^{-1}$, with line-of-sight absorbing columns of $10^{23}-10^{24}$ cm$^{-2}$ \\citep[e.g.][]{krolik1999} causing the soft X-ray band to be highly absorbed \\citep[see~e.g.][~and~references~therein]{gallagher2001}. The vast majority of X-ray observations of BALQSOs have been follow$-$ups after their discovery in optical surveys, and most of the \\emph{XMM-Newton} and \\emph{Chandra} observations of BALQSOs, with the exception of LBQS~2212-1759 \\citep{clavel2006} and APM~08279+5255 \\citep{chartas2002}, have been short snapshots. Our knowledge of the X-ray properties of BALQSOs is therefore based upon a small number of the very brightest optically-selected sources; the largest relevant samples to date have been published by \\citet{green2001,punsly2006,gallagher2006}. Recently a number of X-ray selected BALQSOs (XBALQSOs) have also come to light \\citep[Page~et~al.,~in~preparation;][]{barcons2003}. We use the abbreviation XBALQSOs to refer to serendipitously discovered X-ray sources whose BALQSO nature has been revealed by follow$-$up optical spectroscopic observations. It is unknown whether these represent a distinct section of the BALQSO population. Since the X-ray faintness of BALQSOs is probably due to absorption by the wind, the relative X-ray brightness of XBALQSOs could be due to their winds causing less absorption, through either having lower absorbing columns or higher ionisation parameters; XBALQSOs could also conceivably be exceptionally luminous BALQSOs with high dust extinction in the optical. There is also the possibility that their intrinsic continua simply have a higher ratio of X-ray to optical flux. In this paper, we examine the properties of five XBALQSOs discovered in two deep X-ray survey fields: four objects in the 1$^{H}$ survey field and one in the Chandra Deep Field South (CDFS). We begin by comparing the X-ray and optical fluxes, and X-ray to optical spectral indices ($\\alpha_{ox}$) of the sources to those of the largest available uniformly-analysed sample of OBALQSOs observed to date in the X-ray \\citep{gallagher2006}. We then investigate whether the intrinsic extinction in the XBALQSOs differs from that in the \\citet{gallagher2006} sample and from QSOs from the SDSS \\citep{vandenberk2001}. In section~\\ref{dynamics}, we compare the extent of intrinsic UV absorption in the XBALQSOs with that in the \\citet{trump2006} sample of OBALQSOs from the SDSS. Next, in section~\\ref{xray_wind_properties} we obtain estimates of the absorbing columns and ionisation parameters of the X-ray absorbing winds of the XBALQSOs, comparing the results with those of similar analyses of OBALQSOs in the literature. We then provide estimates of the mass-energy budgets of the XBALQSOs and their X-ray absorbing winds. Finally we discuss the nature of these XBALQSOs, and the implications for the population of faint BALQSOs discoverable in deep X-ray surveys. Names, positions and basic parameters of these sources are given in Table~\\ref{source_properties}. Throughout this paper, the objects are referred to by the abbreviations XBQ1-5 (X-ray selected broad absorption line quasars 1 to 5), as identified in Table~\\ref{source_properties}. We use a cosmology with H$_{0}$=70~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_{M}$=0.3, and $\\Omega_{\\Lambda}$=0.7, and all uncertainties are 1$\\sigma$ unless otherwise stated. \\begin{table*} \\centering \\begin{minipage}{120mm} \\caption{Basic properties of the five X-ray selected BALQSOs: IAU name and abbreviated name for use in this paper; RA; Dec; r, offset between the optical and X-ray positions in arcseconds; redshift (see section~\\ref{optical_data}); Galactic extinction using reference pixel in \\citet{schlegel1998}; de-reddened AB $B$ and $K_s$ magnitudes, where the figures for XBQ5 are from the MUSYC survey (\\citealt{gawiser2006}; Taylor et al., in preparation; see section~\\ref{optical_data}); C$_{x}$, number of background-subtracted X-ray counts in the 0.2$-$12~keV range.} \\begin{tabular}{@{}llllll@{}} \\hline Name & & & RA & Dec & r \\\\ \\hline CXOU J014546.9-043031 & & XBQ1 & 01:45:46.88 & -04:30:30.50 & 0.50 \\\\ CXOU J014517.1-042503 & & XBQ2 & 01:45:17.05 & -04:25:03.20 & 0.91 \\\\ CXOU J014553.9-043726 & & XBQ3 & 01:45:53.93 & -04:37:26.20 & 0.74 \\\\ CXOU J014546.5-042450 & & XBQ4 & 01:45:46.54 & -04:24:49.50 & 0.78 \\\\ CXOCDFS J033209.5-274807 & & XBQ5 & 03:32:09.46 & -27:48:06.80 & 0.28\\footnote{Table 2, \\citet{giacconi2002}} \\\\ \\hline \\hline Name & z & E(B-V) & $B$ & $K_s$ & C$_{x}$\\\\ \\hline XBQ1 & 2.63 & 0.023 & 21.8 $\\pm$ 0.1 & 21.1 $\\pm$ 0.2 & 208 \\\\ XBQ2 & 1.793 & 0.023 & 21.4 $\\pm$ 0.1 & 20.7 $\\pm$ 0.2 & 154 \\\\ XBQ3 & 1.40 & 0.022 & 21.5 $\\pm$ 0.1 & 19.6 $\\pm$ 0.2 & 74 \\\\ XBQ4 & 2.64 & 0.024 & 20.1 $\\pm$ 0.1 & 19.5 $\\pm$ 0.2 & 137 \\\\ XBQ5 & 2.82 & 0.009 & 21.33 $\\pm$ 0.05 & 20.00 $\\pm$ 0.06 & 1115 \\\\ \\hline \\end{tabular} \\end{minipage} \\label{source_properties} \\end{table*} ", "conclusions": "\\label{discussion} Our goal in this paper was to investigate the X-ray and optical properties of a group of X-ray selected BALQSOs discovered in deep X-ray survey fields, and in particular to see whether they had been discoverable in such surveys due to having unusual spectral properties. As we showed in section~\\ref{fluxes_alpha_ox}, the XBALQSOs have, on average, slightly higher X-ray to optical flux ratios than the largest available comparable sample of OBALQSOs \\citep{gallagher2006}. This is probably not a result of intrinsic source variability. It is equally possible that all five sources would happen to be especially X-ray faint as X-ray bright when we observed them, and if variability was a major factor, we would expect to see a mixture of high and low $\\alpha_{ox}$ values among the XBALQSOs. We can also immediately discard the possibility that the flatter $\\alpha_{ox}$ is due to the XBALQSOs being intrinsically very luminous BALQSOs which happen to have high rest-frame optical extinction; as we showed in Section~\\ref{extinction_properties}, their $B-Ks$ colours as a function of redshift are entirely consistent with the OBALQSO population. Given the very low space density of luminous QSOs on the sky, it is also highly unlikely that we should have found four such objects serendipitously in a single 30\\arcmin\\ diameter field. If XBALQSOs are brighter than expected in the X-rays due to having less intrinsic X-ray absorption than optically selected sources, the ionised X-ray absorbing winds should have lower N$_{H}$ and/or higher $\\xi$ than those in comparable OBALQSOs. In Fig.~\\ref{xray_properties_avg} we compare total absorbing column and (weighted average, where necessary) ionisation parameter with these quantities in OBALQSOs from the literature, as well as from a sample of lower redshift Seyfert galaxies and QSOs with ionised soft X-ray absorbers which we have studied previously \\citep{blustin2005}. The OBALQSOs were PG1115+080 \\citep{chartas2007}, Q1246-057, SBS 1542+541 \\citep{grupe2003}, and PG2112+059 \\citep{gallagher2004}. We find that, on the contrary, these XBALQSOs have similar or even perhaps greater intrinsic absorption than the OBALQSOs. Testing whether that is a general feature of XBALQSOs would require a larger sample of sources. \\begin{figure} \\includegraphics[width=55mm,angle=-90]{xray_properties_avg_v6.ps} \\caption{A comparison of total Log $N_{H}$ (the sum of $N_{H}$ for the individual phases) and column-weighted average log $\\xi$ for the X-ray selected BALQSOs in our sample (red stars) with four optically-selected BALQSOs from the literature (black circles; see section~\\ref{discussion}) and a sample of low$-$redshift `warm absorber' Seyferts and QSOs (blue squares; \\citealt{blustin2005}).} \\label{xray_properties_avg} \\end{figure} The amount of absorption from the UV-absorbing part of the wind, as measured by the BI, is consistent with the distribution of BI for OBALQSOs in the SDSS, although there is again some indication that they are towards the more highly absorbed end of the distribution. The mass-energy transport of the winds, which we estimated in Section~\\ref{mass_energy_budget}, ranges quite widely. For XBQ5, which has the best-constrained values, mass outflow rate via the X-ray absorbing part of the wind is $\\sim$~0.04$-$10 times the mass accretion rate onto its black hole, and the kinetic luminosity of the wind is $\\sim$~0.002$-$0.3\\% of the bolometric luminosity. A complete picture of the mass-energy budget of the wind would require consideration of the mass-energy output via the UV-absorbing part of the wind, which is beyond the scope of this paper, and taking account of the full ionisation range of the wind, which would require far better data. Regarding the nature of XBALQSOs, we are left with the conclusion that, although they are fundamentally similar to the wider population of OBALQSOs, they may belong to a part of the BALQSO population with an intrinsically higher X-ray to optical flux ratio. This is actually what we would expect for AGN as UV-faint as our XBALQSOs, since, according to the \\citet{vignali2003} relation, for example, AGN with lower restframe 2500~\\AA\\ luminosities should have flatter $\\alpha_{ox}$ indices. The implication of this is that, as hard X-ray surveys get deeper, there should be a lot of faint XBALQSO-type AGN waiting to be discovered. It is interesting to ask whether the XBALQSO $\\alpha_{ox}$ values do in fact exactly follow the \\citet{vignali2003} relation; our XBALQSOs do not span sufficient luminosity space for us to attempt an answer to that question. Investigating this point will require many more sources over a range of luminosities, with sufficient X-ray counts for a reliable estimation of the unabsorbed X-ray flux to be made." }, "0808/0808.1598_arXiv.txt": { "abstract": "We describe a data reduction pipeline for VLBI astrometric observations of pulsars, implemented using the ParselTongue AIPS interface. The pipeline performs calibration (including ionosphere modeling), phase referencing with proper accounting of reference source structure, amplitude corrections for pulsar scintillation, and position fitting to yield the position, proper motion and parallax. The optimal data weighting scheme to minimize the total error budget of a parallax fit, and how this scheme varies with pulsar parameters such as flux density, is also investigated. The robustness of the techniques employed are demonstrated with the presentation of the first results from a two year astrometry program using the Australian Long Baseline Array (LBA). The parallax of PSR J1559--4438 is determined to be $\\pi = 0.384 \\pm 0.081$\\ mas $(1\\sigma)$, resulting in a distance estimate of 2600 pc which is consistent with earlier DM and HI absorption estimates. ", "introduction": "Accurate, model independent distances to pulsars are, like many astronomical distance measurements, highly prized but difficult to obtain. The dispersion measure (DM) of a pulsar indicates the integrated column density of free electrons between the observer and the pulsar, and can be used in conjunction with Galactic electron distribution models to estimate pulsar distances, but this approach suffers from considerable uncertainty when the electron distribution model is poorly understood, such as at high Galactic latitudes. Timing residuals for millisecond pulsars (MSPs) can be used to determine parallax and proper motion \\citep[e.g.][]{hot06}, but only for the very limited subset of pulsars with extremely accurate timing solutions. VLBI astrometry offers a means to directly measure the parallaxes and proper motions of nearby pulsars. By obtaining an independent pulsar distance estimate, a pulsar DM can be used to determine the average electron density along the line of sight; an ensemble of such measurements can be used to improve electron distribution models. Whilst several recent large pulsar parallax programs have significantly increased the number of known VLBI parallaxes \\citep[e.g.][]{cha04}, to date very few VLBI parallaxes have been obtained for southern pulsars, due to the bias of VLBI facilities towards the Northern Hemisphere. As such, models of Galactic electron distributions are more uncertain at far southern declinations. Furthermore, binary pulsars offer the opportunity to make exceedingly precise tests of gravitational theories, since relativistic effects contribute to the observed rate of change of binary period (\\pbdot), which can be measured very precisely in systems with a stable pulsar. However, kinematic effects \\citep{shk70} also contribute to \\pbdot, and can only be accurately subtracted to obtain \\pbdot\\ due to General Relativity (GR) if the pulsar distance and transverse velocity are accurately known. The effect scales with the square of proper motion, and hence is typically largest for nearby pulsars, where VLBI measurements of parallax are most feasible. Since the uncertainty of a DM--based distance estimate is large for any individual pulsar, a reliable estimate of the error on derived GR quantities requires a direct distance determination. Similarly, luminosities for individual pulsars based on DM--inferred distances are questionable, so deriving accurate an accurate luminosity for any individual pulsar generally requires an independent confirmation of distance. In this paper, we describe our target selection policy in \\S\\ref{sec:select}, and the observations undertaken in \\S\\ref{sec:obs}. The data reduction pipeline is described in \\S\\ref{sec:datared}, and we present the results of PSR J1559--4438 in \\S\\ref{sec:results}. We analyse the optimum visibility weighting scheme to use for VLBI astrometry in \\S\\ref{sec:optweight}, and the magnitudes of different sources of systematic errors are investigated in \\S\\ref{sec:errorbudget}. Our conclusions are presented in \\S\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} A pipeline for the reduction of LBA astrometric data has been developed in ParselTongue and verified by calculating the parallax ($\\pi = 0.384 \\pm 0.081$\\ mas) and proper motion ($\\mu_{\\alpha} = 1.52 \\pm 0.14$\\ mas yr$^{-1}$, $\\mu_{\\delta} = 13.15 \\pm 0.05$\\ mas yr$^{-1}$) of PSR J1559--4438. The calculated values are consistent with the DM distance estimate and earlier HI absorption and proper motion studies. Full account has been made of the impact of residual systematic errors on the quality of the astrometric fit. The optimal weighting scheme in the presence of systematic errors and varying thermal errors has been investigated, resulting in the guideline that superior astrometric quality can be obtained for the LBA for typical astrometric observations by using equally weighted, as opposed to sensitivity weighted, visibilities if the target can be detected with S/N $>10$. The completion of this parallax program will result in the publication of 5 more Southern Hemisphere pulsar parallaxes, which will quadruple the number of Southern Hemisphere pulsars with parallaxes determined directly from VLBI astrometry." }, "0808/0808.1284_arXiv.txt": { "abstract": "When double neutron star or neutron star-black hole binaries merge, the final remnant may comprise a central solar-mass black hole surrounded by a $\\sim 0.01-0.1~ M_{\\sun}$ torus. The subsequent evolution of this disc may be responsible for short $\\gamma$-ray bursts (SGRBs). A comparable amount of mass is ejected into eccentric orbits and will eventually fall back to the merger site after $\\sim 0.01$ seconds. In this {\\it Letter}, we investigate analytically the fate of the fallback matter, which may provide a luminous signal long after the disc is exhausted. We find that matter in the eccentric tail returns at a super-Eddington rate and is eventually ($\\gsim 0.1$ sec) unable to cool via neutrino emission and accrete all the way to the black hole. Therefore, contrary to previous claims, our analysis suggests that fallback matter is {\\it not} an efficient source of late time accretion power and is unlikely to cause the late flaring activity observed in SGRB afterglows. The fallback matter rather forms a radiation-driven wind or a bound atmosphere. In both cases, the emitting plasma is very opaque and photons are released with a degraded energy in the X-ray band. We therefore suggest that compact binary mergers could be followed by an ``X-ray renaissance\", as late as several days to weeks after the merger. This might be observed by the next generation of X-ray detectors. ", "introduction": "Close binaries of compact solar-mass objects are expected to form via the evolution of massive star binaries or by dynamical interaction in dense star clusters. Neutron star (NS--NS) binaries have been detected as radio pulsars \\citep[e.g.][]{faulkner05}, and while black hole--neutron star (BH-NS) or double black hole (BH-BH) binaries have not been observed directly, they are predicted by population synthesis models. The compact objects are expected to merge due to gravitational wave emission, with evolutionary scenarios estimating a local rate of NS--NS mergers $10-100$ times higher than for BH-NS and BH-BH systems \\citep[e.g.][]{bel07}. The final remnant for NS--NS and NS--BH coalescence is generally thought to be a BH of a few solar masses surrounded by a $0.01-0.1~ M_{\\sun}$ accreting disc \\citep[e.g.][]{ruffert97,shibata03,ross04,faber06}. The accretion power immediately following the merger is perhaps the ultimate cause of SGRBs \\citep{blinnikov84,eichler89,paczynski91}. At early times ($\\lsim 0.1 -1$ sec), the accreting disc is geometrically thin, effectively cooled by neutrino emission \\citep{popham99}. When the accretion rate drops below $\\sim 0.1 ~\\msec$ --- the exact value depending on the accretion parameter $\\alpha$ and BH spin (Chen \\& Beloborodov 2007; Metzger, Piro \\& Quataert 2008) --- the disc becomes radiatively inefficient and super-Eddington accretion drives a substantial outflow \\citep{metzger08}. During the dynamical phase of the merger, in which the lighter companion is tidally disrupted, a fraction ($\\sim 10^{-2} M_{\\sun}$) of the debris receives enough energy to be ejected from the system while a comparable amount remains bound in eccentric orbits \\citep[e.g.,][]{ross07,faber06} and will eventually return to the disc site: {\\em fallback} matter. This weakly bound matter may give rise to interesting phenomena observable on timescales longer than any viscous timescale of the disc. For example, it has been suggested \\citep{lee07,ross07,metzger08} that it can be responsible for the X-ray flaring, observed in SGRB afterglows on timescales of minutes to hours \\citep[e.g.][]{campana06}. Unfortunately, numerical investigations have not yet been able to follow the long-term ($>$ minutes) evolution of this eccentric tail, because of time-step limitations \\citep{ross07}. In this {\\em Letter}, we investigate analytically the fate of matter falling back onto a recent merger. We argue that energy released during fallback is {\\it not} a promising source of the X-ray flares. The energy liberated during fallback will either lead to a powerful, radiation-driven wind or a more gradually expanding ``breeze'' that could ultimately form a bound cloud around the merged object. In either case, the expanding gas is so opaque that the radiation is trapped in the expanding flow and degraded to low energies before being released in the X-ray band. We therefore suggest that compact binary mergers might be accompanied by delayed X-ray emission. We assess the detectability of this emission when the merger is localized by either a short $\\gamma$-ray burst or a gravitational wave signal. A direct observation of the accretion activity would give us valuable information on how compact-object binaries merge. This {\\it Letter} is organized as follows. We discuss the behavior of the fallback matter in \\S~\\ref{sec:fb}. Then, we consider two possible scenarios for this material as it rebounds: we model a wind in \\S~\\ref{sec:wind} and a bound atmosphere in \\S~\\ref{sec:atmo}. Prospects for detecting the X-ray emission are discussed in \\S~\\ref{sec:detection} and conclusions are drawn in \\S~\\ref{sec:conclu}. ", "conclusions": "\\label{sec:conclu} In this {\\it Letter}, we investigate the possible fate of fallback matter associated with mergers of compact objects, where a disc is formed by disruption of a NS. Matter flung to highly eccentric orbits will eventually come back to the disc at a super-Eddington rate, converting its kinetic energy into heat via shocks, and will be unable to cool by neutrino emission. Contrary to previous claims, we think that this implies that fallback matter {\\it cannot} accrete all the way to the central object and be responsible for the late energy injections observed in GRB afterglows. Rather, the fallback matter is likely to be blown off the disc plane, leading to the formation of a radiation-driven wind or a bound atmosphere. For the wind case, we have analytically calculated the time evolution of the temperature and luminosity at the trapping radius: while the luminosity decreases (eq.~\\ref{eq:lxwind}), the wind photosphere becomes hotter (eq.~\\ref{eq:txwind}). At first, the emission is in the EUV band and absorption will likely prevent us from observing it. After one or two weeks, the emission finally peaks in the soft X-ray band and the wind activity can be observed. The bound cloud is radiation pressure-dominated and emits at the Eddington limit in soft X-rays, if the atmosphere does not extend much further than $10^{8}$ cm. We note that our estimates of luminosities are conservative: factors such as a smaller electron fraction in the ejected plasma and moderate geometrical beaming can substantially increase the expected luminosity. We also discuss detection prospects for this delayed X-ray activity. Our inspection indicates that only in fortuitous circumstances could the X-ray emission be detected with current instruments, while the planned missions (such as Con-X and XEUS) have a better chance \\S~5.1. Then, the main limiting factors will not be the X-ray detector capability, but rather the tool for localizing the merger \\S~5.2. On the one hand, short $\\gamma$-ray bursts can be easily detected and localized in the whole volume where instruments like {\\it Con-X} and {\\it XEUS} can observe the X-ray emission; however, they are estimated to occur at a rate that is $\\sim 100$ times smaller than the rate at which compact binaries merge. On the other hand, the planned advanced gravitational wave interferometers should be able to detect a signal from any such a merger but within cosmic distances smaller than the maximum distance that {\\it Con-X} and {\\it XEUS} can reach. Moreover, X-ray follow up would require better localization precision than currently estimated. The net result is that between a few to a few tens of detections per year are expected by {\\it XEUS} with a follow-up of a short GRB. Assuming sufficiently good localization, re-pointing after a gravitational signal detection can result in $\\sim 4-54$ wind detection per year from NS-NS mergers, for both {\\it Con-X} and {\\it XEUS}. Furthermore for {\\it XEUS}, there is the exciting possibility to observe X-ray emission from BH-NS mergers: $\\sim 1-53$ event per year. The X-ray emission from these sources should also be brighter than from a NS-NS mergers, since the mass of the central BH could be much larger. The above rates, however, should be taken as indicative of upper limits. We have not taken into account selection effects such as background/foreground sources and the fact that not all BH-NS and NS-NS mergers seem to lead to an accreting system \\citep[e.g.,][]{ross05,bel08}. Moreover, in some cases, the X-ray afterglow from the burst could outshine the wind emission. Nevertheless, the possibility to get information on mergers of compact objects from electromagnetic signals remains, and it could bring important understanding of the physics of these systems. Finally, our findings have implications for interpreting late time activity observed in GRB afterglows. We consider unlikely that fallback matter can be held responsible, since most of the mass is blown away. Even if $\\sim 10 \\%$ of this matter can accrete all the way to the hole, it is very unlikely that it could produce the observed flares, which have an energy ($\\sim 10^{49}-10^{46}$ ergs) comparable to that of the prompt emission \\citep[e.g.][]{campana06}. This would require that the eccentric tail is far more massive than the main disc (contrary to what is observed in simulations) or that the efficiency in converting accreted mass to energy is somehow strongly enhanced in the late fallback accretion. These arguments also apply to the late accretion from the main disc, which is highly super-Eddington \\citep{metzger08}. We thus conclude that, in general, standard late time accretion is unlikely to account for the phenomena, like flares and plateaux, observed in GRB afterglows." }, "0808/0808.3138_arXiv.txt": { "abstract": "We present a coherent theoretical framework for computing gravitational lensing effects and redshift-space distortions in an inhomogeneous universe and investigate their impacts on galaxy two-point statistics. Adopting the linearized Friedmann-Lema\\^\\i tre-Robertson-Walker metric, we derive the gravitational lensing and the generalized Sachs-Wolfe effects that include the weak lensing distortion, magnification, and time delay effects, and the redshift-space distortion, Sachs-Wolfe, and integrated Sachs-Wolfe effects, respectively. Based on this framework, we first compute their effects on observed source fluctuations, separating them as two physically distinct origins: the volume effect that involves the change of volume and is always present in galaxy two-point statistics, and the source effect that depends on the intrinsic properties of source populations. Then we identify several terms that are ignored in the standard method, and we compute the observed galaxy two-point statistics, an ensemble average of all the combinations of the intrinsic source fluctuations and the additional contributions from the gravitational lensing and the generalized Sachs-Wolfe effects. This unified treatment of galaxy two-point statistics clarifies the relation of the gravitational lensing and the generalized Sachs-Wolfe effects to the metric perturbations and the underlying matter fluctuations. For near future dark energy surveys, we compute additional contributions to the observed galaxy two-point statistics and analyze their impact on the anisotropic structure. Thorough theoretical modeling of galaxy two-point statistics would be not only necessary to analyze precision measurements from upcoming dark energy surveys, but also provide further discriminatory power in understanding the underlying physical mechanisms. ", "introduction": "\\label{sec:intro} The standard inflationary models with a single inflaton potential predict a nearly perfect Gaussian spectrum of primordial fluctuations \\citep{BASTTU83,STARO82,HAWKI82,GUPI82,MUFEBR92}. Two-point statistics, correlation function in real space and power spectrum in Fourier space, constitutes a complete description of Gaussian random fields, and it has been widely used to understand the physics of the early universe from measurements of the cosmic microwave background and large-scale structure. The recent discovery \\citep{RIFIET98,PEADET99} of the late time acceleration of the universe has spurred extensive investigations of a mysterious energy component with negative pressure, dubbed dark energy. Observationally, upcoming dark energy surveys will measure galaxy two-point statistics with unprecedented precision from millions of galaxies, constraining the expansion history and the spatial curvature of the universe. Consequently, accurate theoretical modeling of galaxy two-point statistics would be crucial to take full advantage of the promise that these future surveys will deliver. In achieving this goal, complications arise notably from the nonlinear evolution of matter and scale-dependence of galaxy bias. In this paper we limit ourselves to the linear bias model \\citep{KAISE84} and study the linear theory predictions and its corrections, considering that recent attention has been paid to measuring galaxy two-point statistics in the linear regime (e.g., \\citep{EIBLET05,TEEIST06,PENIET07,PASCET07}). However, measurement precision is often highest on nonlinear scales, and proper modeling of galaxy bias on nonlinear scales can substantially increase the leverage to constrain the underlying physics (see, e.g., \\citep{JIMOBO98,SELJA00,MAFR00,PESM00,SCSHET01,BEWE02,COSH02}). Further complication arises from the distortion of redshift-space structure by peculiar velocities, which results in anisotropy from otherwise isotropic two-point statistics \\citep{KAISE87,HAMIL92}. The standard practice is to analyze the angle-averaged correlation function or power spectrum, or to construct a linear combination of their multipole components, suppressing the angular dependence of two-point statistics. However, analyzing the full anisotropic structure, though observationally challenging, can utilize additional information that is lost to some degree in the standard practice \\citep{MATSU04,SEEI07,OKMAET08,GACAHU08}. Gravitational lensing, often assumed to be negligible in galaxy two-point statistics, deflects the propagation of light rays, displacing the position of observed galaxies, and it alters the unit area on the sky and magnifies the observed flux, changing the observed number density of galaxies. The former effect on two-point statistics is to convolve it with the power spectrum of the lensing potential, smoothing out the features in galaxy two-point statistics \\citep{SELJA96}. The latter effect, known as the magnification bias \\citep{NARAY89}, is often used to measure the galaxy-matter cross-correlation function from two source populations separated by large line-of-sight distance \\citep{SCMEET05,BLPOET06}. Recent work \\citep{MATSU00,VADOET07,HUGALO07} showed that these effects on galaxy two-point statistics are non-negligible at the level of accuracy adequate for upcoming dark energy surveys. However, it is unclear whether this list of additional contributions on galaxy two-point statistics is exhaustive, and what are the contribution terms that are ignored in the standard method but need to be considered if higher accuracy is dictated by observations. Here we present a coherent theoretical framework for computing gravitational lensing effects and redshift-space distortions, and investigate their impacts on galaxy two-point statistics in an inhomogeneous universe. Our treatment generalizes the early work \\citep{MATSU00} and complements the recent work \\citep{VADOET07,HUGALO07}, providing a unified description of galaxy two-point statistics. However, we emphasize that these effects naturally arise from metric perturbations in our approach, comprising a complete and exhaustive set of additional (linear order) contributions to galaxy two-point statistic. The rest of this paper is organized as follows. In Sec.~\\ref{sec:formalism}, we describe our notation for the Friedmann-Lema{\\^\\i}tre-Robertson-Walker (FLRW) metric and derive the gravitational lensing and the generalized Sachs-Wolfe effects. In Sec.~\\ref{sec:density}, we study their impacts on source galaxy fluctuations and discuss their correspondence to the standard redshift-space distortion and gravitational lensing effect. In Sec.~\\ref{ssec:two}, we derive the observed galaxy two-point statistics in real space and in Fourier space, and we compare the effects of each contribution term on the observed galaxy two-point statistics in Sec.~\\ref{ssec:com}. We conclude in Sec.~\\ref{sec:discuss} with a discussion of the further improvement of our approach. ", "conclusions": "\\eneq From Eq.~(\\ref{eq:mgb}), the magnification bias is defined as \\beeq \\dMB(\\Vang,z,f)=5\\left[p(f)-0.4\\right]\\kappa(\\Vang,z), \\eneq with redshift $z$ being the source redshift of the convergence $\\kappa(\\Vang)$ in Eq.~(\\ref{eq:conv}). Considering $\\varepsilon(z)\\simeq V(z)-V(0)$, we call the volume effect in Eq.~(\\ref{eq:dgSW}) as the redshift-space distortion bias, \\bear \\label{eq:zfive} \\zdist(\\Vang,z)&=&-2~{1+z\\over H\\chi}~\\varepsilon -(1+z)H{d\\over dz} \\left({\\varepsilon\\over H}\\right)-\\varepsilon+\\delta V \\nonumber \\\\ &=&-2~{1+z\\over H\\chi}~\\varepsilon +{1+z\\over H}~\\varepsilon~{dH\\over dz} \\nonumber \\\\ &&-{1+z\\over H}~{\\partial\\varepsilon\\over\\partial\\chi} -\\varepsilon+\\delta V. \\enar Note that the generalized Sachs-Wolfe effect $\\varepsilon(z)$ implicitly depends on the direction $\\Vang$ via the line-of-sight velocity $V(z)=v_\\alpha e^\\alpha=\\Vang\\cdot\\bdv{v}(\\Vang,z)$, but it is independent of the limiting flux $f$, provided that galaxies have no velocity bias (i.e., galaxies and matter follow the same velocity field). Finally, the evolution bias is defined from Eq.~(\\ref{eq:zevo}) as \\beeq \\devo(\\Vang,z,f)=-{1+z\\over z}\\left[\\alpha-\\beta \\left({z\\over z_0}\\right)^\\beta\\right]\\varepsilon, \\label{eq:bevo} \\eneq where the directional dependence comes from $\\varepsilon$ and the evolution coefficients $(\\alpha,\\beta,z_0)$ depend on the galaxy sample selected with the limiting flux $f$. While the evolution bias arising from the difference between $\\bar n(z)$ and $\\bar n(\\zobs)$ was recognized \\citep{KAISE87,HAMIL98,MATSU04}, it has been ignored in the literature. However, we show in Sec.~\\ref{sec:two} that the evolution bias can be significantly enhanced. Last, we want to emphasize that equation~(\\ref{eq:summary}) is gauge-invariant as is written in the conformal Newtonian gauge. \\label{sec:discuss} Galaxy two-point statistics, correlation function in real space and power spectrum in Fourier space, have been extensively used in cosmology to characterize the underlying matter fluctuations. We have presented a coherent theoretical framework based on the linearized Friedmann-Lema{\\^\\i}tre-Robertson-Walker (FLRW) metric for computing the gravitational lensing and the generalized Sachs-Wolfe effects. Within this framework, the metric perturbations are sourced by the underlying matter fluctuations, and they naturally give rise to perturbations in the observable redshift of source galaxies and their angular position on the sky. The time component of the photon geodesic equations can be used to show the former, the generalized Sachs-Wolfe effect \\citep{SAWO67} that generalizes the standard redshift-space distortion by peculiar velocities in a cosmological context, including the Sachs-Wolfe and the integrated Sachs-Wolfe effects. The spatial components of the photon geodesic equations can be used to derive the latter, the gravitational lensing effect that includes the weak lensing distortion, magnification, and time delay effects. This unified treatment provides a complete description of the relation between these seemingly different effects and the underlying matter fluctuations. Furthermore, it becomes transparent in this treatment how the gravitational lensing and the generalized Sachs-Wolfe effects affect the observed fluctuation field of source galaxies. To the linear order in perturbations, we have computed all the additional contributions to the intrinsic source fluctuation, arising from the gravitational lensing and the generalized Sachs-Wolfe effects. We can gain more insight on the impact of these effects by separating them as two physically distinct origins: the volume and the source effects. The former effect that involves the change of volume is independent of source galaxy populations and hence regardless thereof the volume effect is always present in galaxy two-point statistics. By contraries, the latter effect depends on the intrinsic properties of source galaxy populations and may vanish for a certain population. All of these contributions to the intrinsic source fluctuations result in numerous additional auto and cross terms in the observed galaxy two-point statistics, and therefore proper account should be taken into these additional terms in interpreting measurements of galaxy two-point statistics from upcoming dark energy surveys. With the complete list of the contributions of the gravitational lensing and the generalized Sachs-Wolfe effects, separated as two physically distinct origins, we have identified several contributions in the volume effect and one contribution in the source effect, which are ignored in the standard treatment: the evolution bias in the source effect arises from the generalized Sachs-Wolfe effect, when the mean number density of sources changes rapidly in redshift, and its impact on the observed galaxy two-point statistics can be substantially larger than that of the gravitational lensing magnification bias. The ignored contributions in the volume effect are typically of order peculiar velocities and hence they are subdominant, compared to the standard redshift-space distortion effect. However, their impact is comparable to the magnification bias at low redshift. While the cross term of the magnification bias and the intrinsic source fluctuation is more important at low redshift than the contribution of the magnification bias itself in the gravitational lensing effect, further calculations of the additional contributions associated with the volume effect may be needed, if higher accuracy of theoretical modeling is required from observation. We have investigated the impact of the additional contributions to the anisotropic structure of the observed galaxy two-point statistics, after simplifying some of the contributions to the intrinsic source fluctuations. The redshift-space distortion affects the observed galaxy two-point statistics most, imprinting its well-known feature in the anisotropic structure \\citep{KAISE87,STWI95,HAMIL98}. The gravitational lensing effect is small but non-negligible at a percent level, particularly along the line-of-sight separation and at high redshift, since their contribution increases with longer line-of-sight distance to the source galaxies and the clustering amplitude of the intrinsic source fluctuations decreases in redshift. The evolution bias has an angular pattern similar to the redshift-space distortion, but its impact becomes appreciable, only at fairly large transverse separation. While it is challenging to analyze the observed anisotropic structure of galaxy two-point statistics, its full analysis from upcoming dark energy surveys can provide a great opportunity to separately identify each contribution from the gravitational lensing and the generalized Sachs-Wolfe effects, increasing the leverage to understand the underlying physical mechanism. However, we note that constraining the underlying cosmological model will require not only accurate theoretical predictions, but also model fitting to measurements, which results in further distortion in galaxy two-point statistics, known as Alcock-Paczy\\'nski effect \\citep{ALPA79}. Furthermore, our current investigation has focused on the linear theory predictions and its additional contributions: nonlinearity and scale-dependent galaxy bias can affect our results, though its impact is expected to be less than at the percent level around the acoustic scale (see, e.g., \\citep{EIWH04,SEEI05,EISEWH07}). However, additional leverage can be gained by modeling scale-dependent galaxy bias on nonlinear scales \\citep{YOWEET08}." }, "0808/0808.4004_arXiv.txt": { "abstract": "The near future of astrophysics involves many large solid-angle, multi-epoch, multi-band imaging surveys. These surveys will, at their faint limits, have data on large numbers of sources that are too faint to be detected at any individual epoch. Here we show that it is possible to measure in multi-epoch data not only the fluxes and positions, but also the parallaxes and proper motions of sources that are too faint to be detected at any individual epoch. The method involves fitting a model of a moving point source simultaneously to all imaging, taking account of the noise and point-spread function in each image. By this method it is possible to measure the proper motion of a point source with an uncertainty close to the minimum possible uncertainty given the information in the data, which is limited by the point-spread function, the distribution of observation times (epochs), and the total signal-to-noise in the combined data. We demonstrate our technique on multi-epoch Sloan Digital Sky Survey (SDSS) imaging of the SDSS Southern Stripe. We show that with our new technique we can use proper motions to distinguish very red brown dwarfs from very high-redshift quasars in these SDSS data, for objects that are inaccessible to traditional techniques, and with better fidelity than by multi-band imaging alone. We re-discover all 10 known brown dwarfs in our sample and present 9 new candidate brown dwarfs, identified on the basis of significant proper motion. ", "introduction": "There are many multi-epoch imaging surveys in progress or coming up, which will, among other things, deepen our image of the sky and provide information on source variability and proper motions. These surveys include the SDSS Southern Stripe \\citep{sdssdr7}, the Dark Energy Survey, PanSTARRS, LSST, and SNAP. These surveys promise proper-motion precisions for well-detected sources on the order of $\\masperyr$ over large parts of the sky. For context, a typical halo star at a distance of $10~\\kpc$ moving at a transverse heliocentric speed of $100~\\kmpers$ has a proper motion of $2~\\masperyr$, and a typical disk star at $100~\\pc$ and $10~\\kmpers$ has a proper motion of $20~\\masperyr$. These surveys therefore have the capability of revolutionizing our view of the Galaxy and of the Solar neighborhood. In most conceptions of a proper-motion measurement, one imagines measuring the position of a source in each of several images, taken at different times. A linear trajectory is fitted to the positions, relative to some reference frame or set of fixed sources or sources with well measured proper motions. In its most straightforward form, this method only works for sources bright enough to be detected independently at every epoch---or at least most epochs. In a multi-epoch survey like the SDSS Southern Stripe, which has $\\sim 70$ epochs \\citep{sdssdr7}, this limits the sources with measured proper motions to a small subset of all sources detectable in the combined data, since the combined data reach $\\sim 2.3~\\mag$ fainter than any individual epoch; for typical source populations this represents increases in population size by factors of $5$ to $25$ at any given signal-to-noise threshold. In this paper we present a methodology for measuring in multi-epoch imaging the proper motions of sources too faint to detect at any individual epoch. There are several different technical regimes for these faint-source proper-motion measurements. In the ``easy'' regime, the sources of interest move a distance smaller than or comparable to the point-spread function width over the duration of the multi-epoch survey. In this regime, the sources are easy to detect in the co-added image, even without taking account of their proper motions; proper motions can be determined from processing the individual epoch images after detection in the co-added image. There is a ``difficult'' regime in which the sources of interest move substantially more than the width of the point-spread function over the duration of the survey. In this regime, the source will not appear at high significance in the co-added image if it does not appear at high significance at any epoch, because its different appearances in the different individual-epoch images do not overlap. In principle, the difficult regime can be addressed by brute force with large computing resources. In the context of outer Solar-System bodies, brute-force search in the narrow range of expected motions is feasible (for example, \\citealt{bernstein04a, fuentes2008}). In this paper, we consider only the easy regime. \\paragraph{Modeling the data:} The traditional method for measuring a stellar proper motion with a set of images taken at different times is as follows: Detect the star at each observed epoch; measure its centroid (by, for example, finding the peak or first moment of the flux) at each observed epoch; and fit a linear motion to the measured positions and times. This procedure obtains a proper motion, but it puts an unnecessary requirement on the data: that the star be detectable at every epoch. It also puts an unnecessary burden on the data analyst: it requires decision making about detection and centroiding of the stars at each epoch, decisions that matter at low signal-to-noise, or when faced with data issues such as bad pixels or strong variations in noise from pixel to pixel. Our new approach is to \\emph{model} all individual-epoch images simultaneously with a single point source that is permitted to have a non-zero parallax and proper motion. This approach combines the individual-image positional measurement and the determination of the parallax and proper motion, and determines all of these simultaneously by making a statistically ``good'' model of the union of all the data. In any well-understood imaging survey, each image will have a per-pixel noise model, photometric calibration parameters, and a model of the point-spread function. In any sufficiently small patch of the sky, if the foreground-subtracted intensity in that patch is dominated by a small number of point sources, it is possible to make an accurate model of all of the pixels in the data set that contribute signal to that small patch. In this model of the patch, the fluxes, angular positions, parallaxes, and proper motions of the stars in the patch are simply parameter values in the well-fitted models. In other words, we are assuming that it is possible to model the set of pixels (from all of the images) that contribute to the patch with a $6N$-dimensional model that consists of a set of $N$ moving point sources. The proper motions determined by image modeling have several advantages over those determined by the traditional method: They require fewer decisions about measurement techniques (although they do require a good model of the data, including point-spread function); they use all of the information in all of the pixels, not just those pixels involved in traditional centroiding; they gracefully handle missing data due to bad pixels or cosmic rays (assuming the bad pixels have been flagged); they require the investigator to make explicit the assumptions about the physical properties of the image and the noise; they can be made to properly propagate pixel-value uncertainties into parameter uncertainties (in this case, proper motion uncertainties); they are the result of optimization of a well-justified scalar objective function (in this case the likelihood). Most importantly for what follows, they can be determined in data sets in which the stars are not well detected at any individual epoch, but only appear in the \\emph{combination} of the images. In a data set with $\\sim 70$ similar epochs (such as the SDSS Southern Stripe), this corresponds to an increase in the number of available targets by factors of $5$ to $25$ (assuming source populations double to quadruple with each magnitude of depth). Here we propose, build, test, and use an image-modeling system for the determination of stellar proper motions. We show that it can work down to low signal-to-noise ratios and that it makes measurements in real data that fully exploit the information available. We also use it to discover interesting new astrophysical sources. An approximation to the technique used here has been used previously in the Solar System literature \\citep{bernstein04a}. \\paragraph{Proper-motion and parallax uncertainties:} Consider a well-sampled image $i$ with a point-spread function of full width at half maximum $\\fwhm_i$. The signal-to-noise at which the flux of a point source can be measured, $\\sn_i$, is the sum in quadrature of the signal-to-noise contributions from pixels within the point-spread function. A point source measured with signal-to-noise $\\sn_i$ in a single image can be centroided with (RMS) uncertainty $\\sigma_{\\theta,i}$ of \\begin{equation} \\sigma_{\\theta,i} \\approx \\frac{\\fwhm_i}{\\sn_i} \\quad ; \\end{equation} details such as the shape of the point-spread function introduce factors of order unity \\citep{king1983}. If we have $N$ such images spanning some time interval, we might hope to obtain a proper motion estimate with uncertainty $\\sigma_{\\mu}$ limited by the point-spread function, the time interval, and the total signal-to-noise \\begin{equation} \\sntotal^2=\\sum_{\\mathrm{images}\\ i}\\sn_i^2 \\label{eq:sntotal} \\end{equation} in the combination of all the images (we have assumed here that the images $i$ are all independent). The relevant time ``interval'' is not the total time spanned by the data but rather $\\delta_t\\equiv\\sqrt{\\var{t}}$, the standard deviation (root variance) of the times; the best possible proper-motion estimates will have uncertainties \\begin{equation} \\sigma_{\\mu}\\approx\\frac{\\fwhm}{\\delta_t\\,\\sntotal} \\quad , \\label{eq:muerror} \\end{equation} where properly $\\fwhm$ is the square-signal-to-noise weighted mean point-spread function full width at half maximum, and $\\delta_t$ is the square root of the square-signal-to-noise-weighted variance of the times at which the individual epoch images were taken. By a similar argument, we hypothesize that the best possible parallax estimates will have uncertainties \\begin{equation} \\sigma_{\\pi}\\approx\\frac{\\fwhm}{\\delta_\\lambda\\,\\sntotal} \\quad , \\end{equation} where $\\delta_\\lambda$ is the square root of the square-signal-to-noise-weighted variance of the trigonometric functions of the ecliptic longitude $\\lambda$ of the Sun (time of year in angle units): \\begin{equation} \\delta_\\lambda^2\\equiv\\sigma_{\\cos\\lambda}^2+\\sigma_{\\sin\\lambda}^2 \\quad . \\end{equation} Essentially, $\\delta_\\lambda$ describes how well the parallactic ellipse is sampled; an ideal survey for parallax measurements will have $\\delta_\\lambda\\approx 1$. Disk stars move with respect to one another at velocities of $\\sim 30\\,\\kmpers$ \\citep{dehnen98a, hogg05a}, that is, on the same order as the velocity of the Earth around the Sun. In a multi-epoch survey spanning a small number of years (such as the SDSS Southern Stripe), $\\delta_t$ is of order unity, so for disk stars the parallax and proper motion signal-to-noise ratios ought to be comparable in magnitude. However, most surveys sample ecliptic longitude $\\lambda$ poorly, because of season and scheduling constraints; therefore $\\delta_\\lambda$ is usually substantially less than unity, so the signal-to-noise of parallax is smaller than that of proper motion. ", "conclusions": "We have shown that straightforward image modeling permits the measurement of apparent motions, especially the proper motion and parallax of a source in multi-epoch data, even when the source is too faint to be reliably detected or centroided at any individual epoch. The results of this project are not surprising; indeed what is surprising is how rarely the measurements of stellar motions are made by comprehensive data modeling. We demonstrated the technique on real and artificial data. In the process of performing these tests we showed that spectrosopically confirmed quasars and brown dwarfs can be perfectly distinguished with proper motions measured by this technique. Working without proper motions, but with Co-add Catalog sources and a significant amount of near-infrared imaging follow-up, a group has followed up the $z$-only sources most likely to be high-redshift quasars \\citep{chiu08a, jiang08a}. This project, even after infrared imaging, found---after expensive spectroscopic follow-up---that some of the high-redshift quasar candidates selected on the basis of visible and near-infrared imaging are in fact nearby brown dwarfs. We have shown that all of these spectroscopically confirmed brown dwarfs have significantly measured ($>5$~sigma) non-zero proper motions by the technique shown here (and are reported in \\tablename~\\ref{tab:movingsources}). None of the spectroscopically confirmed high-redshift quasars do. Use of this technique could have been used to substantially increase the efficiency of either quasar or brown-dwarf searches in this data set. In performing this demonstration, we have independently identified all 10 known brown dwarfs \\citep{fan00a, geballe02a, hawley02a, berriman03a, knapp04a, chiu08a, metchev08a} in our parent sample, and we have discovered 9 \\emph{new} candidate brown dwarfs, presented in \\tablename~\\ref{tab:movingsources}. Based on our analysis, these objects have a high probability of being brown dwarfs. It would be desirable to separate disk dwarfs from halo dwarfs---the fastest angular movers tend to be halo members (for example, \\citealt{lepine03a})---but the time cadence of the SDSSSS data is such that parallaxes are not measured well. Two of the dwarfs we rediscover---2MASS J010752.42+004156.3 and 2MASS J020742.84+000056.4---have previously measured parallaxes \\citep{vrba04a}; the measurements are consistent with our upper limits. Our tests show that the uncertainty in the proper-motion measurement made by image modeling is consistent with the best possible uncertainties given the angular resolution and photometric sensitivity of the combination of all images in the multi-epoch data set. These tests effectively show that such measurements can be made for objects that are fainter than those available to traditional methods that require source detection at every epoch. In imaging with $N$ equally sensitive epochs, we are able to measure objects that are fainter by $\\Delta m$ magnitudes: \\begin{eqnarray} \\Delta m &=& -\\log_{2.5}(\\sn_i) + \\log_{2.5}(\\sntotal) \\\\ &=& \\log_{2.5}(\\sqrt{N}) \\\\ &\\sim& 0.55\\,\\log N~\\mag \\quad . \\end{eqnarray} This advantage amounts to $1~\\mag$ for surveys with $6$ similar epochs, and $1.6$ to $2.0~\\mag$ in data with $20$ to $40$ epochs (such as the data used here). In the $\\sim 70$ epochs available in SDSS DR7, it reaches $2.3~\\mag$. Several of the high-redshift quasars and brown dwarfs analyzed in this study were only detectable in the combination of all of the multi-epoch images. The depth advantage of image modeling is most dramatic in surveys with very large numbers of epochs, as is expected for LSST. In general the number of interesting sources is a strong function of depth (factors of $2$ to $4$ per magnitude), so the ``reach'' of the image-modeling technique is a strong function of the number of epochs. One limitation of the work presented here is that we used the zero-proper-motion image ``stack'' for source detection and therefore will only have in the candidate list objects with small proper motions. Faint stars and dwarfs with proper motions large enough that they move the width of the PSF between epochs, or some significant fraction of that, are harder to find, because they don't appear in the stack at much higher signal-to-noise than they appear in any individual-epoch image. In future work we hope to address the detection and measurement of these fast-moving but very faint sources. Approximations have been executed in the search for Solar System bodies (for example, \\citealt{bernstein04a}). Certainly a reliable system for discovery in this regime would have a big impact on future surveys like PanSTARRS and LSST." }, "0808/0808.3630_arXiv.txt": { "abstract": "We present results of a periodicity search of 20 intra-day variable optical light curves of the blazar S5 0716$+$714, selected from a database of 102 light curves spanning over three years. We use a wavelet analysis technique along with a randomization test and find strong candidates for nearly periodic variations in eight light curves, with probabilities ranging from 95\\% to $>$99\\%. This is the first good evidence for periodic, or more-precisely, quasi-periodic, components in the optical intra-day variable light curves of any blazar. Such periodic flux changes support the idea that some active galactic nuclei variability, even in blazars, is based on accretion disk fluctuations or oscillations. These intra-day variability time scales are used to estimate that the central black hole of the blazar S5 0716$+$714 has a mass $> 2.5 \\times 10^6$ M$_{\\odot}$. As we did not find any correlations between the flux levels and intra-day variability time scales, it appears that more than one emission mechanism is at work in this blazar. ", "introduction": "Blazars are the subclass of radio-loud active galactic nuclei (AGNs) consisting of BL Lacertae objects (BL Lacs) and flat spectrum radio quasars (FSRQs). BL Lacs show largely featureless optical continua. All blazars exhibit strong flux variability on diverse time scales varying from a few minutes to many years at all accessible wavelengths of the electromagnetic spectrum. For convenience, blazar variability can be broadly divided into three classes, viz., intra-day or intra-night variability, short-term variability and long-term variability. Variations in flux of up to a few tenths of a magnitude over the course of a night or less is variously known as intra-day variability (IDV) (Wagner \\& Witzel 1995), which is the term we adopt here, microvariability (e.g., Miller, Carini \\& Goodrich 1989), or intra-night optical variability (e.g., Gopal-Krishna, Sagar \\& Wiita 1995) . Short- and long-term variabilities can amount to four or five magnitudes and are usually defined to have time scales from weeks to several months and several months to years, respectively. It is widely accepted that the central engines of AGNs fundamentally are comprised of super massive black holes (SMBHs) along with the radiating matter accreting onto them. In the case of radio-loud AGN such as blazars, jets emerge from these central engines. IDV is most likely produced in close proximity to the SMBH (e.g., Miller et al.\\ 1989), although an origin further away within a turbulent jet (e.g., Marscher, Gear \\& Travis 1992) is also possible. Minimum IDV time scales of blazars can be used to place constraints on sizes of the emitting regions, and, if the radiation is indeed emitted close to the center, on the masses of the SMBHs. Detection of periodic or quasi-periodic oscillations in optical IDV light curves of blazars would be strong evidence for the presence of a single dominant orbiting hot-spot on accretion disk, or accretion disk pulsation (e.g., Chakrabarti \\& Wiita 1993; Mangalam \\& Wiita 1993). Hence the search for periodic or quasi-periodic oscillations in the IDV light curves of blazars is of great interest; however, such variations have not yet been well established. The motivation of the present research program is to see if such periodic IDV patterns do indeed exist in the optical light curves of blazars. For the present work, we extracted data from the published literature for optical IDV of the BL Lac object S5 0716$+$714. It is one of the brightest and therefore most well studied BL Lacs with respect to optical IDV. The optical continuum of this blazar is so featureless that many attempts made to determine its redshift have failed. The non-detection of its host galaxy allowed a lower limit to its redshift, $z > 0.3$ (Wagner et al.\\ 1996), to be set and a later non-detection led to a claim that $z > 0.52$ (Sbarufatti, Treves \\& Falomo 2005). However, there has been a very recent claim of a host galaxy detection which produces a ``standard candle'' value of $z = 0.31 \\pm 0.08$ (Nilsson et al.\\ 2008). The duty cycle of the source is essentially unity, which implies that the source is always in at least a moderately active state (Wagner \\& Witzel 1995). In the search for optical variability on diverse time scales in this source, it has been observed intensely over at least the last 15 years (e.g.\\ Wagner \\& Witzel 1995; Wagner et al.\\ 1996; Heidt \\& Wagner 1996; Ghisellini et al.\\ 1997; Sagar et al.\\ 1999; Villata et al.\\ 2000; Qian, Tau \\& Fan 2002; Raiteri et al.\\ 2003; Wu et al.\\ 2005, 2007; Nesci et al.\\ 2005; Stalin et al.\\ 2006; Gu et al.\\ 2006; Montagni et al.\\ 2006; Ostorero et al.\\ 2006; Foschini et al.\\ 2006; Villata et al.\\ 2008; Gupta et al.\\ 2008 and references therein). Five major optical outbursts were reported for this source; they occurred at the beginning of 1995, in late 1997, in the fall of 2001, in March 2004 and in the beginning of 2007 (Raiteri et al.\\ 2003; Foschini et al.\\ 2006; Gupta et al.\\ 2008). These five outbursts indicate a timescale of long-term variability of $\\sim 3.0 \\pm 0.3$ years (e.g., Raiteri et al.\\ 2003). Correlated radio/optical short-term variability measured over a month seemed to reveal some quasi-periodicities for this blazar (Quirrenbach et al.\\ 1991; Wagner et al.\\ 1996). This paper is structured as follows. In \\S 2 we discuss the source of the data we use and we describe the criteria employed to select the light curves for this study. In \\S 3 we describe the wavelet analysis and randomization technique we employ in our analysis. Our results are presented in \\S 4 and we discuss them and draw conclusions in \\S 5. ", "conclusions": "There are two major classes of models for AGN variability, those involving shocks-in-jets and those invoking instabilities on or above accretion disks. The former is expected to dominate in blazars and the latter is most important when jets are absent or weak (e.g., Wagner \\& Witzel 1995; Urry \\& Padovani 1995; Mangalam \\& Wiita 1993). In another jet-based variant it is argued that for low-luminosity AGNs, the accretion disk is radiatively inefficient and any small amplitude variation, even in the low-state of the source, will be due to a weak jet (e.g., Chiaberge et al.\\ 1999, 2006; Capetti et al.\\ 2007, and references therein). This variant may be able to explain optical IDV in blazars. The detection of quasi-periodicity or periodicity on IDV time scales can be most easily explained by the presence of a single dominating hot-spot on the accretion disk (e.g., Mangalam \\& Wiita 1993; Chakrabarti \\& Wiita 1993) or perhaps by pulsational modes in the disk (e.g., Espaillat et al.\\ 2008). Together, the difficulties in obtaining lengthy, high quality, and essentially evenly spaced data for ground based observations of AGN, and the weaknesses of the standard analysis tools (periodograms and Fourier transforms) under these circumstances, has meant that performing good searches for periods or quasi-periods in blazars has been difficult until recently. So it is not surprising that there have been few strong claims of quasi-periodic variations in blazar LCs. Carini et al.\\ (1992) noted a weak peak in the power spectrum of optical IDV observations of OJ 287 at $\\approx$ 32 minutes, but it was not statistically significant. In the same blazar, Carrasco et al.\\ (1985) earlier claimed periodic variations on timescales of tens of minutes but they were not convincing. Earlier excellent optical data of the blazar 0716$+$714 has shown possible quasi-periodicity on the timescales of $\\sim$ 1 day, 4 days and 7 days, and some of these were in apparent synchrony with radio variations, but the time series were too short to provide conclusive evidence (Wagner 1992; Heidt \\& Wagner 1996). Very recently, using X-ray data for the blazar 3C 273, Espaillat et al.\\ (2008) have reported quasi-periodicity on a timescale of 3.3 ks. If these fluctuations represent orbital periods they would imply central BH masses for 3C 273 some 10 to 100 times smaller than independent determinations have found, so Espaillat et al.\\ (2008) favor the hypothesis that the variations they have seen arise from higher-order modes within the disk. To our knowledge, no other members of the blazar class have shown significant harmonic components in the IDV or short-time scale LCs. Here we have used a wavelet plus randomization technique (Linnell Nemec \\& Nemec 1985; O'Shea et al.\\ 2001) to search for optical IDV time scales of the blazar S5 0716$+$714. We selected 20 nights of IDV data that were characterized by long strings of approximately uniformly sampled data with high IDV signals and low noise from an initial database of 102 nights of observations by Montagni et al.\\ (2006). We found high probabilities ($ \\ge 95$\\%) of (at least quasi-) periodic oscillations in eight of these IDV LCs and very high probabilities ($ \\ge 99$\\%) for five of them. This is the first evidence of detection of periodic components in blazar optical IDV LCs. We found quasi-periodic IDV time scales between $\\sim$ 25 and $\\sim$ 73 minutes. These variability timescales lead to nominal BH masses ranging from 2.47 -- 7.35 $\\times$ 10$^{6}$ M$_{\\odot}$ assuming the period arises from a Schwarzschild BH, while for a rapidly rotating Kerr BH these mass estimates rise to 1.57 -- 4.67 $\\times$ 10$^{7}$ M$_{\\odot}$. However, if these variations arise from internal disk modes then the corresponding BH mass can be substantially larger, by factors of $\\sim$10 to over 100. And if they do emerge from an inner portion of the jet where they are induced by disk fluctuations, the masses would be increased by a factor of $\\delta \\sim 20$; however, if they emerge from jets at large distances from the SMBH, then clearly no information on the BH mass can be extracted from these timescales. We did not find any correlation between average flux and IDV time scales, which implies that optical emission from the blazar is probably not governed by a single mechanism. Our results appear to be most consistent with models in which accretion disk variability plays a role in the optical emission of blazars. Nonetheless, it is possible that emission from turbulet jet, which is also precessing or swinging (e.g., Camenzind \\& Krockenberger 1992; Gopal-Krishna \\& Wiita 1992) could also explain these observations." }, "0808/0808.1918_arXiv.txt": { "abstract": "We have modelled the observed color-magnitude diagram (CMD) at one location in M33's outskirts under the framework of a simple chemical evolution scenario which adopts instantaneous and delayed recycling for the nucleosynthetic products of Type II and Ia supernovae. In this scenario, interstellar gas forms stars at a rate modulated by the Kennicutt-Schmidt relation and gas outflow occurs at a rate proportional to the star formation rate (SFR). With this approach, we put broad constraints on the role of gas flows during this region's evolution and compare its [$\\alpha$/Fe] vs.\\ [Fe/H] relation with that of other Local Group systems. We find that models with gas inflow are significantly better than the closed box model at reproducing the observed distribution of stars in the CMD. The best models have a majority of gas inflow taking place in the last 7 Gyr, and relatively little in the last 3 Gyr. These models predict most stars in this region to have [$\\alpha$/Fe] ratios lower than the bulk of the Milky Way's halo. The predictions for the present-day SFR, gas mass, and oxygen abundance compare favorably to independent empirical estimates. Our results paint a picture in which M33's outer disc formed from the protracted inflow of gas over several Gyr with at least half of the total inflow occurring since $z \\sim 1$. ", "introduction": "\\label{sec:intro} The distribution of stars in a color-magnitude diagram (CMD) is a sensitive function of the past and present star formation rate and chemical composition. This fact allows the extraction of the star formation history (SFH) and chemical enrichment history (CEH) of a galaxy by fitting its observed CMD with a model CMD whose SFH and CEH are known. The results of this synthetic CMD fitting method can ultimately yield insights into the processes shaping galaxy evolution, such as gas flows, energetic feedback, satellite accretion, tidal interactions, and dynamical mixing. Nevertheless, this method is only one of many employed in the domains of near-field cosmology and galactic paleontology \\citep{Freeman02}. Another, equally important method, is chemical modelling, whose primary goal is to reproduce the elemental abundance distribution in a galaxy's stars. This method involves tracking the effects of gas flows, star formation, and stellar nucleosynthesis on the abundances and abundance ratios of certain elements. For example, the abundance of the $\\alpha$-elements (O, Ne, Mg, Si, S, Ca, Ti) relative to iron, commonly plotted as [$\\alpha$/Fe] vs.\\ [Fe/H], is an extremely useful tool to discriminate between different models for a galaxy's evolution because the $\\alpha$-elements and iron have different production sites and time-scales. The $\\alpha$-elements are produced mainly in the hydrostatic burning phases and explosive deaths (supernovae Type II, or SNe II) of massive stars ($M \\ga 8\\ M_{\\sun}$), which have short lifetimes ($\\la 10^7$ yr). The majority of iron, on the other hand, is created in supernovae Type Ia (SNe Ia), which are the thermonuclear explosions of carbon-oxygen white dwarfs in binary systems and, therefore, typically occur on a much longer time-scale of $\\sim 1$ Gyr \\citep{Greggio05}. The methods of CMD fitting and chemical modelling are complementary, but they have evolved mostly independently. Some examples in the recent literature which combine the two methods generally use the results of CMD fitting as external inputs to chemical evolution models. For example, \\citet{Aparicio97} derived the first SFH of the Local Group (LG) dwarf galaxy, LGS 3, from CMD fitting and used the chemical evolution equations to see how much gas inflow and outflow was required to match their results. \\citet{Carigi02} computed the chemical evolution of four dwarf spheroidal (dSph) satellites of the Milky Way (MW). As external constraints, these authors used the SFHs of these galaxies derived by \\citet{Hernandez00} from a CMD fitting method. \\citet{Dolphin02} calculated the SFHs of 6 dSph MW satellites using synthetic CMD fitting, and then \\citet{Lanfranchi03,Lanfranchi04} adopted his SFHs as inputs to their chemical evolution models for the same galaxies. Similarly, \\citet{Fenner06} modelled the chemical evolution of the Sculptor dSph using the SFH derived by \\citet{Dolphin05} from a CMD analysis. Lastly, after adopting the SFH derived by \\citet{Wyder01,Wyder03}, \\citet{Carigi06} modelled the evolution of NGC 6822 in a cosmological framework. Conversely, the results of chemical modelling can be used as external inputs to CMD fitting. For example, \\citet[][hereafter PT98]{Pagel98} modelled the chemical evolution of the Large Magellanic Cloud (LMC) and Small Magellanic Cloud (SMC), adopting gas inflow and non-selective galactic winds (i.e., the wind efficiency was the same for all chemical elements). These authors tuned the model parameters to match the observed elemental abundances of clusters and supergiant stars in these systems. Some subsequent studies adopted the LMC's age-metallicity relation (AMR) derived by \\citetalias{Pagel98} to extract its SFH from the CMD and luminosity function \\citep[e.g.,][]{Holtzman99,SmeckerHane02}. All the aforementioned studies combined CMD fitting and chemical modelling in a two-step process, in which the first step was done independently of the second. However, the two steps are inextricably linked because an increase in the star formation rate (SFR) speeds up the chemical enrichment. Gas flows into or out of the system can change this coupling making them important ingredients to include in any model. Therefore, one drawback of using the CMD-derived SFHs as independent constraints on the chemical evolution models is that the SFHs are not necessarily physically self-consistent under the action of processes like nucleosynthesis, galactic winds, and gas accretion, processes which are fundamental to the chemical models themselves. Similarly, using the results of chemical evolution models as independent constraints on CMD fitting is not necessarily self-consistent because the models have a particular form for the SFH, which is the very thing the CMD fitting is supposed to derive. The recent works of \\citet{Ikuta02} and \\citet{Yuk07} represent a significant improvement over the studies mentioned above because they incorporate CMD fitting and chemical modelling simultaneously. \\citet{Ikuta02} computed a few closed box evolution models for several dSph satellites of the MW. They performed a qualitative comparison of their model CMD and [Mg/Fe] vs.\\ [Fe/H] relation with what was observed and found a reasonable agreement, but they had to invoke gas stripping via ram pressure or tidal shocks to reconcile the present day gas fraction of their closed box models ($\\sim 97\\%$) with the observed values ($\\sim 0\\%$). \\citet{Yuk07} improve upon the work of \\citet{Ikuta02} by quantitatively fitting a closed box chemical model to the CMD of IC 1613, a relatively isolated LG dwarf irregular galaxy. Their model SFH and AMR are in good agreement with previous independent determinations based on the canonical CMD fitting method, lending support to both the old and new methods. Moreover, their predicted present-day oxygen abundance is consistent with the observed value. There are many lines of evidence that suggest not all galaxies evolve as closed boxes. As originally hypothesized by \\citet{Larson74} and exemplified by the \\citet{Ikuta02} results, gas outflows could explain the lack of gas in dSphs despite their low metallicities (see also \\citet{Lanfranchi04}). Other evidence for gas outflows includes the abundances of metals in the IGM \\citep{Edge91,Mushotzky97}, the mass-metallicity relation of galaxies at low and high redshift \\citep[e.g.,][]{Garnett02,Tremonti04,Pilyugin04,Erb06}, extended extra-planar HI gas in spirals \\citep{Fraternali04,Oosterloo07}, high velocity clouds (HVCs) around the MW with near solar metallicity \\citep{VanWoerden04,Wakker01,Richter01}, extra-planar optical and X-ray emission around starburst galaxies \\citep{Heckman90,Lehnert96,Martin02,Strickland04}, and velocity shifts of high-ion absorption lines in damped Lyman-$\\alpha$ systems \\citep{Fox07} and Lyman break galaxies at $z \\sim 3 - 4$ \\citep[e.g.,][]{Pettini01,Adelberger03}. For a more detailed review, we refer the interested reader to \\citet{Veilleux05}. Evidence for gas inflows includes the so-called G-dwarf problem, which is the fact that the metallicity distribution function of low mass, long lived stars observed in the Solar vicinity (SV) and in many other galaxies is too narrow and contains too few metal poor stars compared to the closed box model \\citep[e.g.,][]{Tinsley75,RochaPinto96,Seth05,Jorgensen00,Wyse95, Harris02,Koch06,Sarajedini05,Mouhcine05,Worthey05}. Additionally, gas inflow over several Gyr is required to explain many characteristics of the SV, including stellar chemical compositions and the present-day gas mass fraction and SFR \\citep[e.g.,][]{Chiappini01,Portinari98}. Other evidence for inflows includes the kinematics of extra-planar HI gas in some spirals \\citep{Fraternali08}, the low metallicities and large distances of some MW HVCs \\citep{Wakker99,Richter01}, high velocity OVI absorption along various sight-lines through the MW's halo \\citep{Sembach03}, and the prevalence of warps in HI discs \\citep{Bosma91}. We refer the reader to \\citet{Sancisi08} for a review of observational evidence for gas inflow. There are theoretical expectations for gas inflows, as well. Cosmological simulations of galaxy formation in the $\\Lambda$CDM framework predict disc galaxies to form from the accretion of gas after the last major merger \\citep{Abadi03,SommerLarsen03,Governato04,Governato07}. For galaxies with virialized dark halo masses $\\la 10^{12} M_{\\sun}$, the accreted gas is expected to be mostly cold ($T \\sim 10^{4-5} K$). On the other hand, upon entering the haloes of more massive galaxies, most of the accreted gas experiences shock-heating to the virial temperature ($T \\ga 10^6 K$), creating surrounding reservoirs of hot gas \\citep{Keres05,Dekel06}. Thermal instabilities lead to the condensation of cooler clouds, which then rain down on the discs \\citep{Maller04,SommerLarsen06,Kaufmann06}. In some of the most recent N-body simulations, these clouds have properties similar to the HVCs mentioned above \\citep[][]{Peek08}. Furthermore, this idea is supported by the recent discovery of a hot ($T \\sim 10^6 K$) gaseous halo around the quiescent massive spiral, NGC 5746 \\citep{Pedersen06}. This gas is too hot to be heated by SNe in the disc and the disc SNe rate is too low to have created the reservoir through the outflow of disc gas. Finally, theoretical simulations suggest the need for extended accretion of dilute gas to keep discs from being destroyed after a succession of minor mergers with mass ratios of 4:1 and even up to 10:1 \\citep{Bournaud07}. Despite all this evidence, the precise nature and importance of gas flows in the evolution of galaxies, particularly spirals, is still uncertain. To help improve the situation, in the present paper, we develop a chemophotometric CMD fitting method as an extension to the canonical method used in many other studies, but with the goal of examining the role of gas flows in M33's evolution. Following \\citet{Ikuta02}, we solve the chemical evolution equations to obtain a self-consistent SFH and AMR. We improve upon their work by allowing for gas inflow and outflow and by efficiently searching the full volume of parameter space to make a detailed and quantitative fit to the observed CMD. The photometric data we use in this paper and the reduction procedure were presented in \\citet[][hereafter \\citetalias{Barker07a}]{Barker07a}. In summary, three co-linear fields located in projection $\\sim 20 - 30\\arcmin$ southeast of M33's nucleus were observed with the Advanced Camera for Surveys on board the {\\it Hubble Space Telescope}. In \\citet[][hereafter \\citetalias{Barker07b}]{Barker07b}, we computed the SFHs for these fields using the canonical synthetic CMD fitting method with age and metallicity as free parameters. Because the outer two fields may have a non-negligible contribution from M33's halo or thick-disc (see \\citetalias{Barker07b} for a discussion), we restrict ourselves to analyzing the innermost field in the current study. This field has a projected galactic area of $\\sim 0.7\\ \\rm kpc^2$ and it lies at a deprojected radius of $R_{dp} \\sim 6$ disk scale lengths \\citep{Ferguson07} or $\\sim 9$ kpc, assuming a distance of 867 kpc \\citep{Tiede04}, inclination of $56\\degr$, and position angle of $23\\degr$ \\citep{Corbelli97}. At this radius in M33, the azimuthally averaged HI and stellar surface densities are, respectively, $\\sim 3\\ \\rm M_{\\sun}\\ pc^{-2}$ and $\\sim 0.3\\ \\rm M_{\\sun}\\ pc^{-2}$ \\citep{Corbelli97,Corbelli03}. In the next section, we outline the chemical evolution equations, which form the backbone of our models. We describe in \\S \\ref{sec:method} how we link these equations with the synthetic CMD fitting method to build a self-consistent model CMD. In \\S \\ref{sec:chemresults}, we present the results of fitting closed box and inflow/outflow models to the data. We explore the effects of varying the model parameters in \\S \\ref{sec:var}. In \\S \\ref{sec:obs}, we test the model predictions with independent observations and, in \\S \\ref{sec:chemdisc}, we compare the [$\\alpha$/Fe] vs.\\ [Fe/H] relation in M33 with other LG systems. Finally, we summarize our results in \\S \\ref{sec:chemconc}. ", "conclusions": "\\label{sec:chemconc} In the framework of a simple chemical evolution scenario which adopts instantaneous and delayed recycling for the nucleosynthetic products of Type II and Ia SNe, we have modelled the observed CMD in one location in M33's outskirts. In this scenario, interstellar gas forms stars at a rate modulated by the KS relation and gas outflow occurs at a rate proportional to the SFR. Compared to the common CMD fitting method of allowing age and metallicity to be free parameters, this scenario yields a more physically self-consistent SFH with fewer free parameters and makes more predictions that can be tested with independent observations. Moreover, this method puts broad constraints on the role of gas flows and extends the method of chemical fingerprinting to stellar systems, like M33, that are beyond the reach of current high resolution spectrographs. The precise details of our results depend on which stellar evolutionary tracks are used, Padova or Teramo. Nevertheless, the broad trends appear to be robust. We found that, when the star formation efficiency, $\\epsilon$, is constant, the canonical closed box model fails to reproduce the observed distribution of stars in the CMD. Models with an exponentially declining IFR from 14.1 Gyr ago exhibit similar discrepancies, but to a lesser magnitude. Instead, the inflow models which best reproduce the observed CMD have a significant fraction ($\\ga 50\\%$) of gas inflow taking place in the last 7 Gyr and a smaller fraction ($< 10\\%$) taking place within the last 3 Gyr. This leads to a SFH in overall agreement with what we found in \\citetalias{Barker07b}, and suggests the traditional method of synthetic CMD fitting can be physically self-consistent. Allowing $\\epsilon$ to vary with time, as might be expected if the star formation time-scale depends on local ISM properties, significantly improves the closed box model if the initial metallicity is [Fe/H] $\\sim -1.3$, but the resulting present-day gas mass surface density is too high compared to the observed value. A model with a varying $\\epsilon$ and constant, non-zero IFR reproduces the CMD and gas mass about as well as the best constant $\\epsilon$ models. In this case, more inflow can take place in the last 3 Gyr, but there is still a significant fraction ($\\sim 50\\%$) of inflow occurring over the last 7 Gyr. The amount of variation in $\\epsilon$ required by these models, however, could be larger than the intrinsic scatter in the KS relation of other galaxies. We also examined the [$\\alpha$/Fe] vs.\\ [Fe/H] relation, a key diagnostic of the evolutionary history of stellar systems. Like the MW's dwarf satellites, the bulk of stars at this location in M33's outskirts have [$\\alpha$/Fe] ratios lower than the MW's halo field. Stars formed over 8 Gyr ago have $\\langle$[$\\alpha$/Fe]$\\rangle$ $\\sim 0.2 \\pm 0.1$ and stars forming today have $\\langle$[$\\alpha$/Fe]$\\rangle$ $\\sim -0.1 \\pm 0.1$. The mean [$\\alpha$/Fe] ratio of all stars ever formed at this location in M33 was found to be $\\sim 0.1 \\pm 0.1$. From the tests we have conducted, we estimate that the systematic errors of these values are $\\sim \\pm 0.2$ dex, comparable to that of some high resolution spectroscopic measurements of other systems \\citep[e.g.][]{Monaco05,Geisler07}. Our results paint a picture in which M33's outer disc formed from the protracted inflow of gas over several Gyr with at least half of the total inflow occurring since $z \\sim 1$, but relatively little since $z \\sim 0.25$. All the acceptable inflow models have a similar IFR $\\sim 2 \\times 10^{-4}\\ \\rm M_{\\sun}\\ yr^{-1}\\ kpc^{-2}$ averaged over the lifetime of the Universe. This estimate is uncertain by at least a factor of 2 and the IFR could have intrinsic variations of a factor of $\\sim 3$ when averaged over 2 $-$ 3 Gyr time-scales. Nevertheless, it still gives a rough indication of the mean gas IFR that could occur in the outskirts of low mass spiral galaxies. A useful baseline for comparison is the lifetime average IFR in the SV, predicted by \\citet{Colavitti08} to be $\\sim 4 \\times 10^{-3}\\ \\rm\\ M_{\\sun}\\ yr^{-1}\\ kpc^{-2}$. This estimate comes from several chemical evolution models which adopt different shapes for the IFH and which reproduce many local observables, like the abundances of various elements, MDF of long-lived stars, and present-day gas fraction, total mass density, SFR, IFR, and SN rate. That the mean IFR in the SV has been higher than in M33's outer disc is not surprising since the MW is more massive than M33 and the SV is at a much smaller radius (in terms of disc scale lengths) than the M33 field we have studied. \\citet[][]{SommerLarsen03} ran an N-body cosmological simulation and selected 12 dark matter haloes for more detailed simulations which included baryons. Two Milky Way-type spirals, S1 and S2, that resulted from the simulations experienced gas accretion at a mean rate of $\\sim 10^{-3}\\ M_{\\sun} \\rm\\ yr^{-1}\\ kpc^{-2}$ at radii of $6 - 7$ disc scale lengths. This rate is about 5 times larger than the mean IFR we have found at a similar location in M33's outer disc, but S1 and S2 were about 5 times more massive than M33. The total disc averaged accretion rates of the least massive simulated galaxies, with rotation velocities comparable to M33's, were $\\sim 15 - 35\\%$ smaller than those of S1 and S2 at $z = 0$. Therefore, if the {\\it outer} disc accretion rate scales with a galaxy's rotation velocity in the same way as the {\\it total} disc averaged rate, and if this scaling holds for all redshifts, then our results are consistent with the simulations of \\citet[][]{SommerLarsen03}. What is the source and nature of the gas inflow? If it occurs in the disc plane, it could be caused by spiral density waves, viscosity in the disc gas, or gas with lower angular momentum falling onto the disc at larger radii and flowing toward M33's nucleus \\citep{Lacey85b,Lin87,Bertin96b,Portinari00,Roskar08}. Gas flowing from above the disc plane could come from the condensation of a hot halo corona as described in \\S \\ref{sec:intro}. However, such a process is expected to be more efficient in a massive spiral like the MW than in a late-type spiral like M33 \\citep{Dekel06}. The condensed clouds predicted by numerical simulations to fall onto a disc galaxy have properties similar to the MW's HVC population \\citep{Peek08}. Recent surveys of HI emission around M31 and M33 have revealed an analogous HVC population around these galaxies \\citep{Thilker04,Braun04,Grossi08}. \\citet{Thilker04} detected 20 clouds within 50 kpc of M31's disc and the maps presented by \\citet{Grossi08} show many within $\\sim 20$ kpc of M33 with an inferred total mass $\\geq 5 \\times 10^7\\ M_{\\sun}$. The M31/33 clouds appear to have a mean mass $\\sim 10^5\\ M_{\\sun}$ and size $\\sim 1$ kpc \\citep{Westmeier05,Grossi08}. The lifetime integrated inflow mass of $\\sim 10^6\\ M_{\\sun}$ in the M33 region we have studied requires the accretion of $\\sim 10$ such clouds. Extrapolation to M33's entire disc is risky given how little we know about its evolution and the population of clouds around M33. In addition to condensed halo gas, the clouds could also be low-mass dark matter subhaloes that never formed stars or the tidal debris from recent mergers and interactions \\citep{Thilker04,Grossi08}. In the future, we plan to apply the techniques presented in this paper to other regions of M33 and other galaxies. This approach can provide insights on the origin and amount of accreted gas and coupling (or lack thereof) between the baryonic and dark matter accretion histories of these systems." }, "0808/0808.3295_arXiv.txt": { "abstract": "Five planets are known to orbit the star 55 Cancri. The recently-discovered planet {\\it f} at 0.78 AU (Fischer \\etal 2008) is located at the inner edge of a previously-identified stable zone that separates the three close-in planets from planet {\\it d} at 5.9 AU. Here we map the stability of the orbital space between planets {\\it f} and {\\it d} using a suite of n-body integrations that include an additional, yet-to-be-discovered planet $g$ with a radial velocity amplitude of 5 $m \\, s^{-1}$ (planet mass = 0.5-1.2 Saturn masses). We find a large stable zone extending from 0.9 to 3.8 AU at eccentricities below 0.4. For each system we quantify the probability of detecting planets $b-f$ on their current orbits given perturbations from hypothetical planet $g$, in order to further constrain the mass and orbit of an additional planet. We find that large perturbations are associated with specific mean motion resonances (MMRs) with planets $f$ and $d$. We show that two MMRs, 3f:1g (the 1:3 MMR between planets $g$ and $f$) and 4g:1d cannot contain a planet $g$. The 2f:1g MMR is unlikely to contain a planet more massive than $\\sim 20 \\mearth$. The 3g:1d and 5g:2d MMRs could contain a resonant planet but the resonant location is strongly confined. The 3f:2g, 2g:1d and 3g:2d MMRs exert a stabilizing influence and could contain a resonant planet. Furthermore, we show that the stable zone may in fact contain 2-3 additional planets, if they are $\\sim 50 \\mearth$ each. Finally, we show that any planets exterior to planet {\\it d} must reside beyond 10 AU. ", "introduction": "In a remarkable study, Fischer \\etal (2008) have measured the orbits of five planets orbiting the the star 55 Cancri, the most planets of any exoplanet system to date. The system contains two strongly-interacting, near-resonant giant planets at 0.115 and 0.24 AU (Butler \\etal 1997; Marcy \\etal 2002), a 'hot Neptune' at 0.038 AU (McArthur \\etal 2004), a Jupiter analog at 5.9 AU (Marcy \\etal 2002) and a newly-discovered sub-Saturn-mass planet at 0.78 AU (Fischer \\etal 2008). Table 1 lists Fischer \\etal's self-consistent dynamical fit of the orbits of the five known planets in 55 Cancri. The fast-paced nature of exoplanet discoveries can lead to interesting interactions between theory and observation. Prior to the discovery of planet 55 Cancri {\\it f}, several groups had mapped out the region between planets $c$ and $d$ to determine the most likely location of additional planets. Most studies used massless test particles to probe the stable zone (Barnes \\& Raymond 2004 -- hereafter BR04; Jones, Underwood \\& Sleep 2005; Rivera \\& Haghighipour 2007). Test particles are good proxies for small, Earth-sized planets because they simply react to the ambient gravitational field. However, they are not good substitutes for fully-interacting, real planets. Thus, Raymond \\& Barnes (2005; hereafter RB05) mapped out this zone using Saturn-mass test planets. The stable zone from BR04 and RB05 extended from 0.7 AU to 3.2-3.4 AU, a region that includes the star's habitable zone (Raymond, Barnes \\& Kaib 2006). The planet 55 Cnc $f$ was discovered by Fischer \\etal at the inner edge of that stable zone. The ``Packed Planetary Systems'' (PPS) hypothesis asserts that if a zone exists in which massive planets are dynamically stable, then that zone is likely to contain a massive planet (BR04, RB05, Raymond \\etal 2006; Barnes, Godziewski \\& Raymond 2008). Although the idea behind the PPS hypothesis is not new (see, for instance, Laskar 1996), the large number of planetary systems being discovered around other stars allows PPS to be tested directly. Indeed, the $\\sim$ 1.4 Saturn mass planet HD 74156 $d$ recently discovered by Bean \\etal (2008) was located in the stable zone mapped out in BR04 and RB05, and with the approximate mass predicted by RB05 (Barnes \\etal 2008). In addition, most of the first-discovered planetary systems are now known to be packed (Barnes \\etal 2008), as well as $\\sim$85\\% of the known two-planet systems (Barnes \\& Greenberg 2007). The fact that 55 Cancri $f$ lies within the stable zone identified in previous work (BR04; RB05) also supports PPS, especially since planets $e$ through $c$ are packed, i.e. no additional planets could exist between them. Several other planet predictions have been made and remain to be confirmed or refuted (see Barnes \\etal 2008) -- the most concrete outstanding prediction is for the system HD 38529 (see RB05). Mean motion resonances (MMRs) are of great interest because they constrain theories of planet formation. Models of convergent migration in gaseous protoplanetary disks predict that planets should almost always be found in low-order MMRs and with low-amplitude resonant libration (Snellgrove \\etal 2001; Lee \\& Peale 2002). This may even have been the case for the giant planets in our Solar System (Morbidelli \\etal 2007). On the other hand, planet-planet scattering can produce pairs of resonant planets in $\\sim 5\\%$ of unstable systems, but with large-amplitude libration and often in higher-order MMRs (Raymond \\etal 2008). Thus, understanding the frequency and character of MMRs in planetary systems is central to planet formation theory. In the context of PPS, 55 Cancri is an important system as it contains many planets, but still appears to have a gap large enough to support more planets. Therefore, PPS makes a clear prediction that another planet must exist between known planets $f$ and $d$. In this paper we add massive hypothetical planets to the system identified by Fischer \\etal (2008) to determine which physical and orbital properties could still permit a stable planetary system. We focus our search on the ``new'' stable zone between planets $f$ and $d$. We also show that certain dynamically stable configurations are unlikely to contain a planet because the large eccentricity oscillations induced in the known planets significantly reduce the probability of Fischer \\etal having detected the known planets on their identified orbits, to within the observational errors. The orbital regions that perturb the known planets most strongly correlate with specific dynamical resonances, such that we can put meaningful constraints on the masses of planets in those resonances. Finally, we also use test particle simulations to map out the region of stability for additional planets beyond planet $d$, in the distant reaches of the planetary system. ", "conclusions": "We have mapped out the region in 55 Cancri where an additional planet $g$ might exist. There is a broad region of stability between known planets $f$ and $d$ that could contain a $\\sim$Saturn-mass planet (Fig.~\\ref{fig:ae}). Since observations rule out a very massive planet, our simulations suggest that the region could easily support two or possibly even three additional planets. In addition, one or more outer planets could be present in the system beyond about 10 AU. However, such distant planets would not be detectable for many years. We examined eight mean motion resonances in detail (see Table 2). For two of these, 3f:1g (i.e., the 1:3 MMR between planet $f$ and hypothetical planet $g$) and 4g:1d, there was no stable region that exhibited regular libration of resonant arguments. Therefore, these resonances can not contain planets in the mass range that we explored. Given the very low FTD values, the 2f:1g MMR is unlikely to contain a resonant planet more massive than $\\sim 20 \\mearth$. Two other MMRs, 3g:1d and 5g:2d, may contain a stable, high-FTD resonant planet but the location of the MMRs is constrained to a very small region of ($a_g,e_g$) space which is surrounded by a chaotic region. Finally, three MMRs, 3f:2g, 2g:1d, and 3g:2d, have a stabilizing influence and may contain planets near or even across the collision line with planet $f$ or $d$. Each of these MMRs contains broad regions of stable libration of resonant angles, although the locations of low-FTD libration can vary with $M_g$. We can therefore only weakly constrain the presence of an additional planet in one of these resonances. The region between planets $f$ and $d$ contains many MMRs which display a wide range of behavior. In addition to stable and unstable resonances, the behavior of resonant arguments is also diverse. In some regions we would expect all resonant angles to librate regularly, but in others only some librate. In two instances, planet $g$ could be in the apsidal corotation resonance (Michtchenko \\& Beauge 2003; Ferraz-Mello \\etal 2003): for large $M_g$ in the 2g:1d MMR at the $g-d$ collision line (see Fig.~\\ref{fig:libd21}), or in 5g:2d MMR (Fig.~\\ref{fig:resd52}). Moreover, we also see cases of ``asymmetric'' libration in which the equilibrium angle is neither 0$^\\circ$ or 180$^\\circ$ (see Fig.~\\ref{fig:evold31}). Even if there are no additional planets in the $f-d$ gap, there could be an asteroid belt in which this diverse and exotic dynamical behavior is on display. 55 Cancri is a critical test of the ``Packed Planetary Systems'' (PPS) hypothesis, which asserts that any large contiguous stable region should contain a planet (BR04; RB05; Raymond \\etal 2006; Barnes \\etal 2008). To date, two planets have been discovered in the three stable zones mapped out by BR04 and RB05 (in HD 74156 and 55 Cnc). Given the width of the stable zone between planets $f$ and $d$, PPS indicates that at least one, and possibly two or three, more planet(s) should exist in 55 Cancri. We look forward to further observations of the system that may find such planets, or perhaps show evidence of their absence. Our results may be used to guide observers searching for planet $g$ and beyond." }, "0808/0808.0543_arXiv.txt": { "abstract": "The temperatures of electrons and ions in the post-shock accretion region of a magnetic cataclysmic variable (mCV) will be equal at sufficiently high mass flow rates or for sufficiently weak magnetic fields. At lower mass flow rates or in stronger magnetic fields, efficient cyclotron cooling will cool the electrons faster than the electrons can cool the ions and a two-temperature flow will result. Here we investigate the differences in polarized radiation expected from mCV post-shock accretion columns modeled with one- and two-temperature hydrodynamics. In an mCV model with one accretion region, a magnetic field \\raisebox{-0.2em}{$\\stackrel{>}{\\sim}$} 30 MG and a specific mass flow rate of $\\sim$0.5 g cm$^{-2}$ s$^{-1}$, along with a relatively generic geometric orientation of the system, we find that in the ultraviolet either a single linear polarization pulse per binary orbit or two pulses per binary orbit can be expected, depending on the accretion column hydrodynamic structure (one- or two-temperature) modeled. Under conditions where the physical flow is two-temperature, one pulse per orbit is predicted from a single accretion region where a one-temperature model predicts two pulses. The intensity light curves show similar pulse behavior but there is very little difference between the circular polarization predictions of one- and two-temperature models. Such discrepancies indicate that it is important to model some aspect of two-temperature flow in indirect imaging procedures, like Stokes imaging, especially at the edges of extended accretion regions, were the specific mass flow is low, and especially for ultraviolet data. ", "introduction": "Magnetic cataclysmic variables (mCVs), consisting of the classes known as polars and intermediate polars, are composed of a Roche-lobe-filling M type main sequence star in orbit about a magnetic white dwarf (see \\citet{warner95} for a review). Mass is lost through the inner Lagrangian point, $L_{1}$, and flows toward the magnetosphere of the white dwarf either predominately in a stream (polars) or after forming a truncated accretion disk circulating around the white dwarf (intermediate polars). In either case, the ionized gas follows magnetic field lines to the surface of the white dwarf after the gas reaches the magnetosphere where the magnetic pressure exceeds the gas ram pressure. Upon reaching the white dwarf surface the gas will be essentially at ``free fall'', with highly supersonic velocities. The abrupt stop of the radial inflow near the surface of the white dwarf leads to the formation of a shock, which heats the inflowing material \\citep{fabian76,king79,lamb79,wu00}. The hot subsonic post-shock flow settles gradually onto the white dwarf, and cools via emitting bremsstrahlung X-rays and optical/infra-red cyclotron radiation. The hydrodynamic structure of the post-shock settling flow is determined by radiative and particle energy processes, which are essentially characterized by the bremsstrahlung cooling time $t_{\\mbox{\\small br}}$, the cyclotron cooling time $t_{\\mbox{\\small cy}}$, the electron-ion energy-exchange time $t_{\\mbox{\\small ei}}$, the electron-electron collisional time $t_{\\mbox{\\small ee}}$, and the ion-ion collisional time $t_{\\mbox{\\small ii}}$ \\citep{king79,lamb79,imamura96,saxton05}. For weakly magnetic systems (with $B \\sim 10^6~$G or weaker) with accretion luminosities $10^{31}-10^{33}$ erg cm$^{-2}$ s$^{-1}$, typical of mCVs, $t_{\\mbox{\\small cy}} > t_{\\mbox{\\small br}} > t_{\\mbox{\\small ei}} > t_{\\mbox{\\small ee}}$. As the strength of the magnetic field increases to $B \\: \\raisebox{-0.2em}{$\\stackrel{>}{\\sim}$} \\: 10^7$~G, cyclotron cooling may dominate bremsstrahlung cooling, $t_{\\mbox{\\small cy}} < t_{\\mbox{\\small br}}$. For sufficiently strong magnetic fields and low specific accretion rates, $t_{\\mbox{\\small cy}}$ is so short ($t_{\\mbox{\\small cy}} < t_{\\mbox{\\small ei}}$) that collisions between electrons and ions cannot maintain an equal temperature between the two types of particles. The accretion flow is therefore in a two-fluid regime which requires a two-temperature (2T) hydrodynamic description. A strong magnetic field can also result in a situation where cyclotron radiative loss is so rapid ($t_{\\mbox{\\small cy}} < t_{\\mbox{\\small ee}}$) that electron-electron collisions are not efficient enough to maintain a Maxwellian distribution. In the extreme situation where $t_{\\mbox{\\small cy}} < t_{\\mbox{\\small ii}}$, the accretion flow is no longer hydrodynamic. Previous calculations of cyclotron radiation from the post-shock settling flow in mCVs have either assumed a uniform density and temperature (\\citet{chanmugam81}; \\citet{meggitt82};\\\\ \\citet{barrett84};\\\\ \\citet{wickramasinghe85};\\\\ \\citet{wu88,wu89}) or a one-temperature (1T) structure (\\citet{wu90,wu92}; \\citet{potter02}). However, detailed 1D calculations of the hydrodynamic structure of a post-shock accretion column that self-consistently include cyclotron and bremsstrahlung cooling clearly show that a 2T structure is to be expected in many physical situations relevant to mCVs (\\citet{imamura96}; \\citet{woelk96}; \\citet{saxton01}; \\citet{wu03}; \\citet{saxton05}). Here we have computed and compared the cyclotron radiation from a cylindrical post-shock accretion column, with a uniform cylindrical radial structure, assuming both a 1T hydrodynamic structure and a 2T hydrodynamic structure. The resulting cyclotron spectra for a grid of three white dwarf masses (0.5, 0.7 and 1.0 $M_{\\odot}$), three magnetic field strengths (10, 30 and 50 MG) and two mass flow rates ($10^{16}$ and $10^{14}$ g s$^{-1}$) were computed for various viewing inclination angles. For each case, using the computed viewing-angle dependent cyclotron spectra, Johnson bandpass \\citep{johnson65,bessell90} filtered light curves over an orbital period were computed for a mCV with an orbital inclination of 45$^{\\circ}$ and with the given accretion column at a co-latitude of 30$^{\\circ}$. This work is organized as follows. In \\S \\ref{methods} we outline the hydrodynamic formulation used to determine the density and temperature structure of the post-shock flows and the radiative transfer through the ionized accreting gas. In \\S \\ref{results} we present the results of the polarized radiative transfer calculations, and in \\S \\ref{discussion} we examine the differences between the spectro-polarization properties of the emission from 1T and 2T accretion flows and discuss their implications. Concluding remarks are made in \\S \\ref{conclusion}. ", "conclusions": "Assuming a one-temperature hydrodynamic post-shock accretion column as the source for polarized radiation in models of magnetic cataclysmic variables can lead to erroneous predictions for the radiation when the cyclotron cooling efficiency is greater than the electron-ion energy exchange efficiency. This effect shows up at the lower mass flow rate modeled here ($\\dot{m} = 0.5$ g cm$^{-2}$s$^{-1}$) at higher cyclotron harmonics for the fairly generic white dwarf masses of 0.7 and 1.0 $M_{\\odot}$ and magnetic fields of 30 MG or greater. So the interpretation of light curve data obtained at higher frequencies, in the ultraviolet, needs to take into account the effect of two-temperature flow on the production of polarized radiation. In particular, the number of linear polarization pulses observed in the ultraviolet can be misinterpreted if a one-temperature accretion column flow is assumed." }, "0808/0808.3226_arXiv.txt": { "abstract": "We present a spectral line survey of the C-rich envelope \\object{CIT\\,6} in the $\\lambda$ 2\\,mm and 1.3\\,mm bands carried out with the Arizona Radio Observatory (ARO) 12\\,m telescope and the Heinrich Hertz Submillimeter Telescope (SMT). The observations cover the frequency ranges of 131--160\\,GHz, 219--244\\,GHz, and 252--268\\,GHz with typical sensitivity limit of $T_R<10$\\,mK. A total of 74 individual emission features are detected, of which 69 are identified to arise from 21 molecular species and isotopologues, with 5 faint lines remaining unidentified. Two new molecules (C$_4$H and CH$_3$CN) and seven new isotopologues (C$^{17}$O, $^{29}$SiC$_2$, $^{29}$SiO, $^{30}$SiO, $^{13}$CS, C$^{33}$S, and C$^{34}$S) are detected in this object for the first time. The column densities, excitation temperatures, and fractional abundances of the detected molecules are determined using rotation diagram analysis. Comparison of the spectra of \\object{CIT\\,6} to that of \\object{IRC+10216} suggests that the spectral properties of \\object{CIT\\,6} are generally consistent with those of \\object{IRC+10216}. For most of the molecular species, the intensity ratios of the lines detected in the two objects are in good agreement with each other. Nevertheless, there is evidence suggesting enhanced emission from CN and HC$_3$N and depleted emission from HCN, SiS, and C$_4$H in \\object{CIT\\,6}. Based on their far-IR spectra, we find that \\object{CIT\\,6} probably has a lower dust-to-molecular gas ratio than \\object{IRC+10216}. To investigate the chemical evolution of evolved stars, we compare the molecular abundances in the AGB envelopes \\object{CIT\\,6} and \\object{IRC+10216} and those in the bright proto-planetary nebula \\object{CRL\\,618}. The implication on the circumstellar chemistry is discussed. ", "introduction": "The late stages of stellar evolution from the asymptotic giant branch (AGB) to planetary nebulae (PN) are now recognized as an active period of chemical synthesis of molecules. The detection and analysis of millimeter wave molecular emission lines are fundamental to the understanding of the physical conditions and chemical processes leading to chemical synthesis. Due to the rapid evolution of the star, the changing physical conditions, including dust, stellar winds, shock waves, UV emission and X-rays from the central star, etc. play different roles in circumstellar chemistry. This leads to corresponding different circumstellar chemical compositions in different evolutionary stages. The envelopes around C-rich stars, with their enhanced carbon abundance, provide a perfect cradle for molecule formation. Hitherto, more than 60 molecular species have been detected in C-star envelopes \\citep{glassgold96,olo97,cernicharo00,ziu07}, most of which were discovered through their rotational lines at millimeter wavelengths. Recent improvement in telescope design and receiver performance enable us to detect new molecular emission with a higher sensitivity, and thus with the possibility of shedding new light on circumstellar chemistry. The most frequently investigated C-star envelope is \\object{IRC+10216}, which is one of the richest molecular sources in the sky. Several molecular line surveys have been presented for this object \\citep[see][and the references therein]{cernicharo00,he08}, which was found to harbor extremely abundant carbon chain and metal-containing molecules. \\object{IRC+10216} has been frequently used as a standard reference for the chemical compositions of late-type stars. This inevitably invites the issue whether \\object{IRC+10216} is a chemically unique late-type star. To settle this question, we require systematic surveys of molecular line emission from other C-star envelopes. In the present study, we report a spectral-line survey of the C-star envelope \\object{CIT\\,6} at millimeter wavelengths. This allows us to compare the similarity and difference in the chemical compositions between the two C-star envelopes. \\object{CIT\\,6} (\\object{RW\\,LMi, GL\\,1403, IRC+30219, IRAS\\,10131+3049}) was first discovered during the Caltech 2-$\\mu$m sky survey and was among the 14 very red infrared-bright optical-faint sources found \\citep{ulrich66}. \\object{CIT\\,6} is characterized by its very low color temperature, implying that the star is surrounded by a very thick dust envelope and has been identified as a long-period variable with a period of about 628 days \\citep{alksnis95}. From the period-luminosity relation, \\citet{cohen96} estimated the distance of \\object{CIT\\,6} to be $400\\pm50$\\,pc, which is slightly more distant than \\object{IRC+10216}, which has a distance of between $\\sim120$\\,pc \\citep{groen98} and $\\sim150$\\,pc \\citep{gue99}. \\object{CIT\\,6} is believed to be more evolved and has a lower mass loss rate compared to \\object{IRC+10216} \\citep[see][e.g.]{fukasaku94}. The large polarization found in the visible and infrared wavelengths implies that the distribution of circumstellar material around the star is asymmetric \\citep[][]{kruszewski68,dyck71}. Multi-wavelength imaging observations have been performed to study the structure of the nebula around \\object{CIT\\,6}. The optical images obtained by the Hubble Space Telescope (HST) and the near-IR images of \\object{CIT\\,6} obtained by the Keck-I telescope have revealed a bipolar dust envelope and an elongated component with time-variable asymmetry \\citep{monnier00}. Several scattering arcs were revealed by the HST-NICMOS imaging polarimetry \\citep{schmidt02}. These arcs are nearly concentric and extend to large stellar radii. Mid-IR images of \\object{CIT\\,6} were obtained by \\citet{lagadec05} using the ESO 3.6-m telescope. A cometary-like feature was revealed in their 9.7\\,$\\mu $m image. There have been several observations of molecular lines in \\object{CIT\\,6} at millimeter wavelength. \\citet{henkel85} reported observations of a few molecular lines in \\object{CIT\\,6} and \\object{IRC+10216} between 18 and 150\\,GHz. They found that relative abundances of observed molecules in the two sources have no significant differences. Using the Nobeyama 45\\,m radio telescope, \\citet{fukasaku94} observed a few transitions in the frequency ranges between 39--47\\,GHz and 85--91GHz in a sample of evolved stars including \\object{CIT\\,6} and found that the abundance of HNC increases with the evolutionary stage of the stars. \\citet{bujarrabal94} presented observations of 10 molecular transitions in C-rich and O-rich circumstellar envelopes including \\object{CIT\\,6} with the IRAM 30\\,m radio telescope at 1.3\\,mm, 2\\,mm, and 3\\,mm windows. A recent molecular line survey was presented by \\citet{woods03} using the SEST 15\\,m and Onsala 20\\,m telescopes. They found that \\object{CIT\\,6} stands out from the other C-rich envelopes due to its high CN/HCN ratio and low HNC/HCN ratio. To date, the molecular species positively detected in \\object{CIT\\,6} at millimeter wavelengths include CO, $^{13}$CO, CN, $^{13}$CN, CS, SiO, SiS, $^{29}$Si$^{32}$S, C$_{2}$H, SiC$_{2}$, HCN, H$^{13}$CN, HNC, C$_{3}$N, HC$_{3}$N, HC$^{13}$CCN, HCC$^{13}$CN, and HC$_{5}$N. In this paper, we present the first systematical line survey of \\object{CIT\\,6} at the 2\\,mm and 1.3\\,mm windows, using the Arizona Radio Observatory (ARO) 12\\,m telescope and the Heinrich Hertz Submillimeter Telescope (SMT). The observations are described in Sec.~2. In Sect.~3 we present the identifications and abundance calculations of the detected molecular species. In Sect.~4 we discuss the implication of our findings on circumstellar chemistry. The conclusions are given in Sect.~5. ", "conclusions": "The presence of rich molecular species around evolved stars provides an opportunity to study the evolution of chemistry in circumstellar envelopes, which have been widely suggested as one of the main sources of organic compounds in space. As part of our project of investigating circumstellar chemistry, this paper reports a spectral line survey of the carbon-rich envelope \\object{CIT\\,6}, covering the frequency range between 131--160, 219--244, and 252--268\\,GHz with a high sensitivity. A total of 74 lines are reported in the survey. We identify 69 lines belonging to 21 different molecular species and isotopologues, most of which are carbon-bearing species. The new species include two carbon-chain molecules, C$_4$H and CH$_3$CN, and seven C, O, S, and Si isotopologues. Several new transitions from known species have been detected for the first time in this object. The species with the largest number of detected emission lines in our survey is SiC$_2$, which has 19 lines. It is followed by HC$_3$N, with 7 lines. We find that the line profiles for some molecular species have different shapes, suggesting that the chemical structure is asymmetric in the envelope. A comprehensive 3D photochemistry model is required to account for the line intensities and profiles in \\object{CIT\\,6}. The excitation temperatures, column densities and abundances of the detected molecules are determined through rotation-diagram analysis. The spectra of \\object{CIT\\,6} are characterized by a large CN/HCN abundance ratio. Our results suggest that there is evidence for the photodissociation of HCN and SiS and the formation of CN and HC$_3$N in the evolved AGB envelope. The strong SiO and CS emission may suggest that depletion onto dust grains and destruction by shocks are insignificant in this object. An abundance comparison with the PPN \\object{CRL\\,618} implies to a rapid chemical evolution after a star leaves the AGB stage. In order to investigate whether the molecular environment of \\object{IRC+10216} is intrinsically unique, we systematically compare its spectra with those of \\object{CIT\\,6}. According to the comparison, we find that the molecular species can be classified into three groups, a) for most of the species, the intensity ratios of individual lines in the two objects are in good agreement with each other; b) the emission from HC$_3$N and CN may be enhanced in \\object{CIT\\,6}; c) the emission from SiS, HCN, and C$_4$H may be depleted in \\object{CIT\\,6}. The differences of the line-intensity ratios in the two objects are probably a consequence of chemical evolution with the exception of C$_4$H, for which a high abundance in \\object{IRC+10216} cannot be explained by photochemical models. The ISO LWS spectra show that \\object{CIT\\,6} has a lower continuum-to-line ratio than \\object{IRC+10216}, suggesting that the latter might have a larger dust-to-molecular gas ratio. Using the same telescope settings, we also obtained the spectra of the AGB stars \\object{IRC+10216} and \\object{CRL\\,3068}, the PPN \\object{CRL\\,2688}, and the young PN \\object{NGC\\,7207}. A detailed study of the chemical compositions in different evolutionary stages will be published in a separate paper." }, "0808/0808.1824_arXiv.txt": { "abstract": "The Einstein-Aether theory provides a simple, dynamical mechanism for breaking Lorentz invariance. It does so within a generally covariant context and may emerge from quantum effects in more fundamental theories. The theory leads to a preferred frame and can have distinct experimental signatures. In this letter, we perform a comprehensive study of the cosmological effects of the Einstein-Aether theory and use observational data to constrain it. Allied to previously determined consistency and experimental constraints, we find that an Einstein-Aether universe can fit experimental data over a wide range of its parameter space, but requires a specific rescaling of the other cosmological densities. ", "introduction": " ", "conclusions": "" }, "0808/0808.3683_arXiv.txt": { "abstract": "We present a new method to estimate the average star formation rate per unit stellar mass (SSFR) of a stacked population of galaxies. We combine the spectra of $600-1000$ galaxies with similar stellar masses and parameterise the star formation history of this stacked population using a set of exponentially declining functions. The strength of the Hydrogen Balmer absorption line series in the rest-frame wavelength range $3750-4150$\\AA\\ is used to constrain the SSFR by comparing with a library of models generated using the BC03 stellar population code. Our method, based on a principal component analysis (PCA), can be applied in a consistent way to spectra drawn from local galaxy surveys and from surveys at $z \\sim 1$, and is only weakly influenced by attenuation due to dust. We apply our method to galaxy samples drawn from SDSS and DEEP2 to study mass-dependent growth of galaxies from $z \\sim 1$ to $z \\sim 0$. We find that, (1) high mass galaxies have lower SSFRs than low mass galaxies; (2) the average SSFR has decreased from $z=1$ to $z=0$ by a factor of $\\sim 3-4$, independent of galaxy mass. Additionally, at $z \\sim 1$ our average SSFRs are a factor of $2-2.5$ lower than those derived from multi-wavelength photometry using similar datasets. We then compute the average time (in units of the Hubble time, $t_{\\rm H}(z)$) needed by galaxies of a given mass to form their stars at their current rate. At both $z=0$ and at $z=1$, this timescale decreases strongly with stellar mass from values close to unity for galaxies with masses $\\sim 10^{10} M_{\\odot}$, to more than ten for galaxies more massive than $ 10^{11} M_{\\odot}$. Our results are in good agreement with models in which AGN feedback is more efficient at preventing gas from cooling and forming stars in high mass galaxies. ", "introduction": "\\label{sec:intro} Over the last decade, there have been a large number of photometric and spectroscopic surveys designed to study the formation and evolution of galaxies. One major conclusion of these studies has been that the epoch when massive galaxies formed most of their stellar mass is significantly earlier than that for low mass galaxies \\citep{Heavens04, Thomas05}. This phenomenon, popularly known as ``down-sizing'', at first sight seems at odds with the predictions of the hierarchical CDM model, in which dark matter halos of all masses grow through merging and accretion right up to the present day. The only way to reconcile the observations with the theory, is to postulate that the growth of the most massive galaxies is much slower than the growth of their surrounding halos. The most natural way to achieve this is by invoking feedback processes, which prevent gas from cooling, condensing and forming stars in massive halos \\citep{Bower06, Croton06, DeLucia06, Guo08}. A variety of different feedback mechanisms have been included in numerical and semi-analytic models, for example feedback from quasars and radio AGN \\citep{Silk98, Hopkins06, wang06, Wang07}, supernova heating \\citep{Cole00, Benson03, Stringer08} and heating by infalling substructure \\citep{Dekel08}. A physical understanding of feedback is still lacking, however, and it is not understood which, if any, of the proposed mechanisms are most important in regulating the growth of galaxies. It is likely that each mechanism will come into play at a different mass scale and cosmological epoch. By quantifying in detail how galaxies of different masses grow as a function of time, we hope to clarify how different galaxies form their stars and how this is influenced by feedback processes. There have been many attempts to measure the star formation rates of galaxy populations from the present day out to redshifts greater than 5 \\citep[see for example][]{Brinchmann04, Bauer05, Feulner05, Noeske07a}. For nearby galaxies, the H$\\alpha$ line provides the most reliable estimate of SFR, since it directly measures the number of ionizing photons from massive stars. A reasonably reliable correction for dust extinction can be made provided one also measures the Balmer decrement H$\\alpha$/H$\\beta$ accounting accurately for stellar absorption. Beyond redshifts of $\\sim 0.4$, the H$\\alpha$ redshifts out of the optical part of the spectrum and is no longer accessible. [O {\\sc ii}] equivalent widths, ultraviolet (UV)-optical spectral energy distribution (SED) fitting and infrared photometry are commonly used to estimate the star formation rate. Each of these indicators is subject to different disadvantages. The [O {\\sc ii}] equivalent width is strongly affected by dust and by metallicity. The UV luminosities of galaxies are also strongly affected by dust, while the infrared only provides a direct measure of SFR if one has measurements across the thermal peak of the spectrum. This has not been the case for many of the Spitzer surveys aimed at quantifying star formation in high-redshift galaxies. All the indicators discussed above may also be contaminated by AGN emission if there is an actively accreting black hole in the galaxy. In this paper, we develop an approach to estimate the amount of recent star formation experienced by a {\\em population} of galaxies. Our method is based on the Balmer absorption lines located in the rest-frame wavelength range $3750-4150$\\AA\\ of the galaxy spectrum. The advantages of our method are the following: (1) The Balmer absorption lines are weakly influenced by dust attenuation and AGN contamination compared with the indicators discussed above. (2) The wavelength range spanned by the Balmer absorption lines is accessible out to redshifts greater than 1, even in optical spectra. This means that the method can be applied in a consistent manner to both low and high redshift ($z \\sim 1$) samples. (3) By stacking a large number of galaxies, we estimate SFRs for a {\\em complete sample} of galaxies in a given stellar mass range, including those galaxies that are forming stars weakly or not at all. These galaxies are often excluded when SFRs are measured for individual objects. We apply our method to a large sample of galaxy spectra from the Sloan Digital Sky Survey (SDSS) and the DEEP2 redshift survey to study the evolution of galaxies from $z=1$ to $z=0$. This redshift interval accounts for roughly half the age of the universe. Our study thus addresses the final stage of galaxy build-up in the Universe. This paper is arranged as follows. In \\S2, we introduce the SDSS and DEEP2 samples used in our studies. The method to estimate the amount of recent star formation in our galaxies is developed in \\S3. We apply the method to the SDSS and DEEP2 samples, and present our results in \\S4 and \\S5, respectively. A discussion of the results is given in \\S6. \\S7 contains the summary of the paper. We use the cosmological parameters $H_0=70~{\\rm km~s^{-1}~Mpc^{-1}}$, $\\Omega_{\\rm M}=0.3$ and $\\Omega_{\\Lambda}=0.7$ throughout this paper. ", "conclusions": "In this section, we translate SSFR into a dimensionless star formation activity parameter (see Dav\\'e 2008), defined as \\begin{equation} \\alpha_{sf} \\equiv \\frac{1}{\\rm SSFR}\\frac{1}{t_{\\rm H}(z)-1 \\rm Gyr}. \\end{equation} Physically, this can roughly represent the fraction of the Hubble time (minus a Gyr) that a galaxy needs to have formed its stars at its current rate in order to produce its current stellar mass\\footnote{We note that $\\alpha_{sf}$ is not simply the fractional Hubble time the galaxy takes to form its stars, because stellar mass is returned to the ISM, a fact which is not accounted for in this simple model.}. A Gyr is subtracted in order to take account of the fact that dark matter halos massive enough to host galaxies with reasonably high star formation rates take about 1 Gyr to assemble in a $\\Lambda$CDM Universe \\citep{Dave08}. A value of $\\alpha_{sf}$ of 1 indicates that the galaxies could feasibly have formed all their stellar mass by forming stars continuously at the rate now observed. If $\\alpha_{sf}$ is greater than 1, their past average SFR must have been greater than their current SFR for the stellar mass of the galaxy to have formed within a Hubble time. In Figure 10, we show our estimate of this characteristic timescale $\\alpha_{sf}$ as a function of stellar mass at both high (red asterisks) and low (black diamonds) redshifts. We remind the reader that the low-$z$, highest mass bin is uncertain due to aperture correction complications in the SDSS survey. From this plot, we find that $\\alpha_{sf}$ is a strongly increasing function of $M_*$; the timescale increases from values close to a Hubble time for galaxies with $10^{10} M_{\\odot}$ to more than an order of magnitude larger than the Hubble time for galaxies more massive than $10^{11}M_{\\odot}$. We find that, at fixed stellar mass, $\\log \\alpha_{sf}$ evolves only slightly with redshift, decreasing by around 0.2dex between $z\\sim0$ and $z\\sim1$. \\begin{figure} \\bc \\hspace{-0.6cm} \\resizebox{8.5cm}{!}{\\includegraphics{fig/f10.ps}}\\\\% \\caption{Star formation activity parameter $\\alpha_{sf}$ as a function of stellar mass. As in previous plots, black diamonds and red asterisks represent SDSS and DEEP2 results, respectively. Blue squares are the results derived from the DEEP2 mock catalogue, they have been displaced in the x-axis by plus 0.05dex to make the comparison clear.} \\ec \\end{figure} \\subsection{Comparison with galaxy formation models} The blue squares in Figure 10 (displaced in the x-axis by plus 0.05dex to make the comparison clear) show the characteristic timescale of star formation in the mock DEEP2 universe created from the Millennium run SAM as described in Section 5.2. A key result of this paper is that the model is entirely consistent with observational results at $z\\sim1$, both in amplitude and slope. More generally, in theoretical models of galaxy formation, at a fixed stellar mass, $\\alpha_{sf}$ is predicted to remain constant out to redshifts greater than $2$ \\citep[Figure 2 of][]{Dave08}. This latter result is robust to methodology, both semi-analytic and smooth-particle-hydrodynamic (SPH) simulations agree. Specifically, between a redshift of 1 and 0 the models predict changes in $\\log\\alpha_{sf}$ of less than 0.1dex. Our observational results indicate a slightly larger increase of 0.2dex. Further studies using larger surveys will be required to confirm this small amount of evolution. The amplitude of $\\alpha_{sf}$ and its variation with stellar mass depend on simulation methodology. The near unity $\\alpha_{sf}$ at both $z\\sim0$ and $z\\sim1$ for galaxies with $M_* \\la 10^{10.5}M_{\\odot}$ suggests that galaxy mass growth in this mass and redshift range may be dominated by smooth and steady cold mode accretion, as implemented in all current models of galaxy formation. The strong increase of $\\alpha_{sf}$ with $M_*$ that we observe at both redshift 1 and in the local Universe (Figure 10), is an expression of the phenomenon of `downsizing': massive galaxies have apparently completed most of their star formation at higher redshifts than low mass systems. In many current models of galaxy formation, the explanation of this behaviour is ``AGN feedback''. More massive galaxies are more likely to host massive black holes which have the capability of producing more energy. Of equal importance, these massive galaxies are hosted by larger halos, where the AGN energy can be well coupled to the material that would otherwise cool and fuel star formation in the galaxies. As a result, AGN feedback through heating of the interstellar and intergalactic gas is more efficient in massive galaxies. In summary, the observed amplitude and evolution of $\\alpha_{sf}$ as presented in this paper, provide firm constraints for all galaxy formation models. \\subsection{Comparison with previous results} As we have shown, our results appear to be in good general agreement with cosmological galaxy formation models. We now turn to a comparison with previous results from the literature. As we shall show, there is a significant discrepancy of $0.3-0.4$dex between our $H_{\\rm Balmer}$ derived SSFRs, and those derived primarily from multiwavelength broad band photometry. \\begin{figure} \\bc \\hspace{-0.6cm} \\resizebox{8.5cm}{!}{\\includegraphics{fig/f11.ps}}\\\\% \\caption{SSFR as a function of stellar mass. Black diamonds: aperture-corrected SDSS data. Red asterisks: SSFRs of DEEP2 galaxies. Blue crosses and green squares: SSFRs from Z07 \\& AEGIS respectively.} \\ec \\end{figure} The results from two other studies of the SSFRs of galaxies at $z \\sim 1$ are compared with our results in Figure 11. The blue crosses represent SSFRs derived from the COMBO-17 sample with Spitzer 24$\\mu$m and GALEX data (Zheng et al. 2007, hereafter Z07). These authors use the measured UV+IR luminosities to derive SFRs \\citep{Bell05}. They take account of the contribution from galaxies that are not individually detected at 24$\\mu$m by stacking their images. The green squares show results calculated by us for a sample of galaxies selected from the All-Wavelength Extended Groth Strip International Survey (AEGIS). The galaxies are a subsample of those used to create the composite spectra in this paper. This sample is not exactly the same as presented in Noeske et~al. (2007, hereafter N07) and \\citet{Noeske07b}, as only star-forming galaxies were included in their analysis. Instead, we have averaged the SSFRs over all the galaxies in each stellar mass bin (i.e. both star-forming and quiescent ones), using the same weighting factors that we used in our spectral stacking analysis. Aside from this difference in sample, the method used to derive SFRs for the AEGIS galaxies is the same as used by N07. The method used by N07 to derive SFRs is slightly different to that employed by Z07, in that information from emission lines in the spectra of the galaxies is utilized if it is available. For galaxies with $f_{24\\mu m}>60\\mu$Jy and strong emission lines, the total SFR is derived from a combination of the IR measurements and from DEEP2 emission lines (H$\\alpha$, H$\\beta$, or [O {\\sc ii}]$\\lambda$3727, depending on $z$) with no extinction correction. SFRs are derived from extinction-corrected emission lines only for blue galaxies with strong emission lines and no detectable 24 micron emission. Red galaxies with weak emission lines, but no 24 micron detections, are considered star-forming and SFRs are derived from emission lines, assuming the same extinction corrections as for normal star-forming galaxies. The dashed lines in Figure 11 are a linear fit to the data points for mass bins with $\\log M_{*}/M_\\odot \\ge 10$. As can be seen from this plot, the slope of the relation between SSFR and mass that we derive is consistent with the results of AEGIS and Z07. However our method yields a normalisation that is $0.3-0.4$dex lower than AEGIS and Z07. This discrepancy is puzzling, but it does have a number of possible explanations: 1) {\\em Systematic differences in the calibration of different SFR indicators.} Our method is based on stellar absorption line indicators, while the N07 and Z07 results are based on a combination of UV, IR and emission lines. We have demonstrated that our results do agree with the SSFRs derived by B04 from emission lines at low redshift, which provides confidence that there is no significant discrepancy between our method and the standard calibration of extinction-corrected H$\\alpha$ to derive SFR. We are unable to carry out the same test at $z \\sim 1$, because H$\\alpha$ is redshifted out of the spectral range covered by most galaxy surveys. Systematic differences in estimated stellar mass do occur when different population synthesis models are used due to differing mass-to-light ratios in the models. However, the same BC03 models have been used in the comparison to the AEGIS results. 2) {\\em Obscured AGN.} There has been no attempt to remove obscured AGN from the two $z=1$ galaxy samples with which we compare. Both the N07 and Z07 analyses make use of the 24 micron Spitzer passband as a star formation indicator. At $z=1$, this corresponds to a rest-frame wavelength of 12 micron, which is very close in wavelength to where emission from a dusty torus would become very significant \\citep[see e.g.][]{Daddi07}. In addition, AGN emission could well be contaminating some of the optical emission lines used to estimate SFR. By contrast, the $H_{\\rm balmer}$ index originates from stellar atmospheres and is not expected to be contaminated from emission from an obscured AGN. On the other hand, the vast majority of DEEP2 24 micron sources and line--emitting galaxies have line ratios indicating star formation and not AGN \\citep{Weiner07}. If the AGN were highly obscured, they could contribute at 24 micron and not show up in optical lines, but to make up a difference of 0.3dex one would have to assign half of the 24$\\mu$m emission at z=1 to obscured AGN, which would appear quite extreme (B.~Weiner, private communication). 3) {\\em Evolution of IMF with redshift.} The SFR indicators used by N07 and Z07 trace O and B stars, whereas our SFR indicator is the Balmer series and is mainly influenced by A stars. If the IMF changes with redshift such that more massive stars form at higher $z$ \\citep[e.g.][]{Dave08, Dokkum08}, then a discrepancy between the two methods may not be apparent in the analysis of the low redshift samples, but may become more pronounced at higher $z$. 4) {\\em Aperture effects.} As we have discussed, the SSFRs we estimate for the $z=1$ galaxies may be biased somewhat low because the long-slit spectra preferentially sample the inner bulge of the galaxy. However, 95\\% of the DEEP2 galaxies (at all redshifts) with a line measurement have $r_{eff} < 0.95^{\\pp}$ as measured in the CFHT imaging \\citep{Weiner07}. Thus the 1$^{\\pp}$ slit covers a large fraction of the galaxy. Additionally the seeing mixes the light into the slit to a much greater degree in DEEP2 than it does for the SDSS fibers. Thus the star formation gradients of $z=1$ galaxies would have to be extremely strong for this to make a factor $2-3$ difference to the SSFR estimated for the galaxy population as a whole. As can be seen from Figure 8, the correction for aperture effects in the low $z$ sample is only a factor of 2 on average, even in the case where the fibre only samples 25\\% of the total light. In this paper, we are not able make a definitive conclusion with regard to the possibilities listed above. It is clear that there are many inherent uncertainties in estimating SFRs in galaxies and that more work is needed before the factors of $2-3$ offsets that we see between the different methods and the models can be understood in detail." }, "0808/0808.2685_arXiv.txt": { "abstract": "In this paper un-binned statistical tools for analyzing the cosmic ray energy spectrum are developed and illustrated with a simulated data set. The methods are designed to extract accurate and precise model parameter estimators in the presence of statistical and systematic energy errors. Two robust methods are used to test for the presence of flux suppression at the highest energies: the Tail-Power statistic and a likelihood ratio test. Both tests give evidence of flux suppression in the simulated data. The tools presented can be generalized for use on any astrophysical data set where the power-law assumption is relevant and can be used to aid observational design. ", "introduction": "\\label{sec:Intro} The observation of suppression in the flux of the highest energy cosmic rays (CRs) has been of central interest to astro-particle physics since the prediction of the GZK-effect\\cite{refG,refZK} in 1966. Most recently both the Auger\\cite{Yamamoto:2007xj} and the HiRes\\cite{Abbasi:2007sv} detectors have released results favoring the observation of flux suppression at a $6\\sigma$ and $5\\sigma$ level of confidence, respectively. With this in mind, we describe a set of statistical tools designed to extract the most accurate and precise information concerning the flux of the highest energy cosmic rays. By binning the data we can only lose information\\cite{refGolds} (see \\secref{sec:BvUB}) and therefore our statistical tools use an un-binned maximum likelihood approach\\cite{refPDG, refHowell, refNewm, refClauset07} to answer two related statistical questions: {\\it Is there flux suppression at the highest energies?} and, if yes, {\\it What are the characteristic cut-off energy and shape parameters?} In detail we first generate a toy data set using the CRPropa package\\cite{refCRPropa}, as in \\secref{sec:ats}. We then fit this simulated data to the three models described in \\secref{sec:Models}. The un-binned maximum likelihood fit is outlined in \\secref{sec:ID} and methods for incorporating systematic and statistical energy errors are described in \\secref{sec:SysEE} and \\secref{sec:StatEE} respectively. In \\secref{sec:EtF} we describe several statistical tools for hypothesis testing: the Kolmogorov-Smirnov test, the tail power statistic\\cite{refPisa1, refHague, Yamamoto:2007xj}, and a likelihood ratio test\\cite{refHagueicrc1217}. Though we cast our discussion in terms of cosmic ray energies, it is worth noting that these tools can be applied to any astrophysical data set where deviations from the power-law hypothesis are relevant, e.g. the galaxy correlation function\\cite{refZehavi} or gamma ray astronomy\\cite{refSchroedter}. ", "conclusions": "\\label{sec:SaC} In this paper we describe a set of statistical tools designed to extract the most accurate and precise information about the flux of the highest energy cosmic rays. We show how to use the un-binned likelihood method described in \\secref{sec:ID} to fit a data set to the three model distributions described in \\secref{sec:Models}. Techniques for incorporating the systematic and statistical errors associated with a real CR detector into the likelihood method are described in \\secref{sec:SysEE} and \\secref{sec:StatEE} respectively. In \\secref{sec:EtF} we describe $p$-values useful for extracting information about flux suppression. We show in \\secref{sec:TP} and \\secref{sec:MD} how an experimenter might use an {\\it a priori} estimate of the cut-off energy to maximize an observational setup for detecting flux suppression. The collection of these statistical tools are the primary result of this paper. To answer the questions posed in the introduction for a given data set we suggest the following steps: \\begin{enumerate} \\item Estimate the best fit parameters $\\hat{\\theta}$ of the model; \\label{step:BF} \\begin{enumerate} \\item The estimates $\\tgH$, $\\tebH$ or $\\tehH$ and $\\tdH$ or $\\twH$ are determined via the likelihood \\equref{equ:lnQ},\\label{step:BFa} \\item The estimate of the minimum energy $\\teminH$ is that which minimizes the Kolmogorov distance $\\DKS$ (see \\secref{sec:GoF}).\\label{step:BFb} \\end{enumerate} \\item Shift the energies up and down according to the systematic uncertainty described in \\secref{sec:SysEE} and repeat step (\\ref{step:BF}). The resulting shift in parameter estimates gives the systematic uncertainty of those estimates. \\label{step:SysE} \\item Obtain the model parameter estimates using the methods in \\secref{sec:StatEE} to incorporate the statistical error of each event energy. \\label{step:StatE} \\item Test the model hypothesis; \\label{step:GoF} \\begin{enumerate} \\item The absolute goodness of fit for any of the models can be evaluated using $\\pKS$ in \\secref{sec:GoF},\\label{step:GoFa} \\item The Tail-Power statistic $\\pTP$ can be used to reject the single power-law hypothesis (nearly independently of the spectral index estimate, see \\secref{sec:TP})\\label{step:GoFb} \\item The single power-law may be rejected in favor of a specific alternative model using $\\pR$, here we study the double and Fermi power-law distributions (see \\secref{sec:MD}).\\label{step:GoFc} \\end{enumerate} \\end{enumerate} The best estimates for the {\\it characteristic cut-off energy and shape parameters}, determined via steps (\\ref{step:BF}), (\\ref{step:SysE}) and (\\ref{step:StatE}), are $\\tebH$ or $\\tehH$ and $\\tdH$ or $\\twH$ respectively. The presence of {\\it flux suppression at the highest energies} can be evaluated using step (\\ref{step:GoF}). By applying these methods to the toy Monte-Carlo set of CRPropa events we illustrate in \\secref{sec:SumToy} how the procedure may be implemented on an actual CR detector, i.e. a detector with systematic and statistical event energies. Suppression in the tail is clear in \\figref{fig:cdf} and \\figref{fig:rescdf}; the tail power statistic is $4.6\\sigma$ and the $p$-value for the double (Fermi) power-law is $\\lg p_{\\text{DP}} = -2.7$ ($\\lg p_{\\text{FP}} = -1.9$). The methods are sufficient and robust. Indeed, many of them have been applied by the Auger collaboration which reports suppression with $6\\sigma$ confidence\\cite{Yamamoto:2007xj}. These tools serve as a basis for further investigation of the CR spectrum such as evidence for more detailed spectral information. They can be applied to any data set, astrophysical or otherwise, to provide information both about data already collected and help to optimize future observations for detecting tail suppression. \\newpage \\appendix" }, "0808/0808.0363_arXiv.txt": { "abstract": "We report on the results of systematic infrared 2.5--5 $\\mu$m spectroscopy of 45 nearby ultraluminous infrared galaxies (ULIRGs) at $z <$ 0.3 using the IRC infrared spectrograph onboard the AKARI satellite. This paper investigates whether the luminosities of these ULIRGs are dominated by starburst activity, or alternatively, whether optically elusive buried active galactic nuclei (AGNs) are energetically important. Our criteria include the strengths of the 3.3 $\\mu$m polycyclic aromatic hydrocarbon (PAH) emission features and the optical depths of absorption features at 3.1 $\\mu$m due to ice-covered dust grains and at 3.4 $\\mu$m from bare carbonaceous dust grains. Because of the AKARI IRC's spectroscopic capability in the full 2.5--5 $\\mu$m wavelength range, unaffected by Earth's atmosphere, we can apply this energy diagnostic method to ULIRGs at $z >$ 0.15. We estimate the intrinsic luminosities of extended (several kpc), modestly obscured (A$_{\\rm V}$ $<$ 15 mag) starburst activity based on the observed 3.3 $\\mu$m PAH emission luminosities measured in AKARI IRC slitless spectra, and confirm that such starbursts are energetically unimportant in nearby ULIRGs. In roughly half of the observed ULIRGs classified optically as non-Seyferts, we find signatures of luminous energy sources that produce no PAH emission and/or are more centrally concentrated than the surrounding dust. We interpret these energy sources as buried AGNs. The fraction of ULIRGs with detectable buried AGN signatures is higher in ULIRGs classified optically as LINERs than HII-regions, and increases with increasing infrared luminosity. Our overall results support the scenario that luminous buried AGNs are important in many ULIRGs at $z <$ 0.3 classified optically as non-Seyferts, and that the optical undetectability of such buried AGNs occurs merely because of a large amount of nuclear dust, which can make the sightline of even the lowest dust column density opaque to the ionizing radiation of the AGNs. ", "introduction": "Ultraluminous infrared galaxies (ULIRGs) have large luminosities ($L> 10^{12}L_{\\odot}$) that are radiated mostly as infrared dust emission \\citep{san88a,sam96}. This means that very luminous energy sources ($L> 10^{12}L_{\\odot}$) are present hidden behind dust, so that most of the energetic photons from these energy sources are once absorbed by the surrounding dust and the heated dust then emits strongly in the infrared. The ULIRG population dominates the bright end of the luminosity function in the local universe ($z <$ 0.3) \\citep{soi87}. The contribution from ULIRGs to the total infrared radiation density increases rapidly with increasing redshift \\citep{lef05} and becomes important at $z >$ 1.3 \\citep{per05}. Distinguishing whether the dust-obscured energy sources of ULIRGs are dominated by starburst activity (nuclear fusion inside stars), or whether active galactic nuclei (AGNs; active mass accretion onto compact supermassive black holes with M $>$ 10$^7$M$_{\\odot}$) are also energetically important, is closely related to unveiling the history of dust-obscured star formation and supermassive black hole growth in the universe. Identifying the dust-obscured energy sources in ULIRGs is difficult because a large amount of molecular gas and dust is concentrated in the nuclear regions of ULIRGs \\citep{dow98,ink06,ima07b}, and can easily {\\it bury} (= obscure in all directions) the putative AGNs because the emitting regions of AGNs are spatially very compact. Unlike AGNs obscured by torus-shaped dust, which are classified optically as Seyfert 2s, such {\\it buried} AGNs are elusive to conventional optical spectroscopy \\citep{mai03}, and thus very difficult to detect. However, quantitative determination of the energetic importance of buried AGNs is indispensable in understanding the true nature of the ULIRG population. To study such optically elusive buried AGNs in ULIRGs, observing them at wavelengths of low dust extinction is important. Infrared 3--4 $\\mu$m (rest-frame) spectroscopy is one such wavelength. Dust extinction at 3--4 $\\mu$m is as low as that at 5--13 $\\mu$m \\citep{lut96}. More importantly, starburst and AGN emission are distinguishable from spectral shapes based on the following two arguments. (1) A normal starburst galaxy should always show large equivalent width polycyclic aromatic hydrocarbons (PAH) emission at rest-frame 3.3 $\\mu$m, while an AGN shows a PAH-free continuum originating in AGN-heated hot dust emission \\citep{moo86,imd00,idm06}. Hence, if the equivalent width of the 3.3 $\\mu$m PAH emission is substantially smaller than that of starburst-dominated galaxies, then a significant contribution from an AGN to the observed 3--4 $\\mu$m flux is the most natural explanation \\citep{imd00,idm01,im03,idm06,ima06}. (2) In a normal starburst, the stellar energy sources and dust are spatially well mixed (Figure 1a), while in a buried AGN, the energy source (= a compact mass accreting supermassive black hole) is more centrally concentrated than the surrounding dust (Figure 1b). In a normal starburst with mixed dust/source geometry, the optical depths of dust absorption features at 3.1 $\\mu$m by ice-covered dust and at 3.4 $\\mu$m by bare carbonaceous dust cannot exceed certain thresholds, but can be arbitrarily large in a buried AGN \\citep{im03,idm06}. Therefore, detection of strong dust absorption features whose optical depths are substantially larger than the upper limits set by mixed dust/source geometry suggests the buried-AGN-like centrally concentrated energy source geometry \\citep{im03,idm06}. Infrared 3--4 $\\mu$m slit spectroscopy of a large sample of nearby ULIRGs using infrared spectrographs attached to ground-based large telescopes was performed, and signatures of intrinsically luminous buried AGNs were found in a significant fraction of optically non-Seyfert ULIRGs \\citep{imd00,idm01,im03,idm06,ris06,ima06,ima07b,san08}. However, the observed sample was restricted to $z <$ 0.15 because above this redshift, a part of the above-mentioned spectral features pass beyond the Earth's atmospheric window ($L$-band; 2.8--4.1 $\\mu$m), making this 3--4 $\\mu$m energy diagnostic method impossible. Because of the 2.5--5 $\\mu$m spectroscopic capability of the IRC \\citep{ona07,ohy07}, which was mounted onboard the AKARI satellite \\citep{mur07} and thus unaffected by Earth's atmosphere, we can now extend this successful approach to more distant ULIRGs at $z> $ 0.15. As AKARI IRC spectroscopy from space is quite sensitive, given the lack of large background emission from Earth's atmosphere, this extension to higher redshift is feasible in terms of sensitivity. Additionally, the AKARI IRC employs slitless spectroscopy, so that the bulk of the extended starburst emission in host galaxies ($>$ several kpc scale) is covered, unless it is extended more than 1 $\\times$ 1 arcmin$^{2}$ (Onaka et al. 2007). The observed 3.3 $\\mu$m PAH emission luminosities in the AKARI IRC spectra can thus be used to roughly estimate the intrinsic luminosities of modestly obscured (A$_{\\rm V}$ $<$ 15 mag) extended starbursts in the ULIRG host galaxies, which is impossible with ground-based spectroscopy using a narrow ($<$ a few arcsec) slit. Based on observations of several sources, nearby ULIRGs are argued to be energetically dominated by highly obscured {\\it compact} ($<$1 kpc) nuclear cores, with small contributions from {\\it extended}, modestly obscured starbursts in host galaxies \\citep{soi00,fis00}. The AKARI IRC spectra can directly test whether this argument holds for the majority of nearby ULIRGs. In this paper, we present the results of systematic 2.5--5 $\\mu$m slitless spectroscopy of a large number of nearby ULIRGs at $z <$ 0.3 using the AKARI IRC. Throughout this paper, $H_{0}$ $=$ 75 km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{\\rm M}$ = 0.3, and $\\Omega_{\\rm \\Lambda}$ = 0.7 are adopted to be consistent with our previous papers \\citep{idm06,ima06}. The physical scale of 1$''$ is 0.74 kpc at $z =$ 0.040 (the nearest source), 2.44 kpc at $z =$ 0.15, and 3.84 kpc at $z =$ 0.268 (the farthest source). ", "conclusions": "We investigate energy sources of observed ULIRGs based on the 3.3 $\\mu$m PAH emission and 3.1 $\\mu$m and 3.4 $\\mu$m absorption features. \\subsection{Modestly obscured starbursts} In most of the observed ULIRGs, the presence of starbursts is evident from detection of the 3.3 $\\mu$m PAH emission. Since dust extinction at 3.3 $\\mu$m is only $\\sim$1/15 of that in the optical $V$-band ($\\lambda$ = 0.6 $\\mu$m; Rieke \\& Lebofsky 1985; Lutz et al. 1996), the flux attenuation of 3.3 $\\mu$m PAH emission with dust extinction of A$_{\\rm V}$ $<$ 15 mag is less than 1 mag. Thus, the observed 3.3 $\\mu$m PAH emission luminosities can be used to quantitatively derive the intrinsic luminosities of modestly obscured (A$_{\\rm V}$ $<$ 15 mag) starburst activity. The observed 3.3 $\\mu$m PAH to infrared luminosity ratios (L$_{\\rm 3.3PAH}$/L$_{\\rm IR}$) are summarized in column 4 of Table 3. Figures 4a and 4b respectively compare the observed 3.3 $\\mu$m PAH luminosity (L$_{\\rm 3.3PAH}$) and its rest-frame equivalent width (EW$_{\\rm 3.3PAH}$) measured in ground-based slit spectra (abscissa) and in AKARI IRC slitless spectra (ordinate). The abscissa and ordinate of Figure 4a trace nuclear ($<$kpc) and total starburst luminosities, respectively. If {\\it extended} (several kpc) starbursts in the host galaxies of ULIRGs are energetically much more important than modestly obscured {\\it nuclear} starbursts, then both the L$_{\\rm 3.3PAH}$ and EW$_{\\rm 3.3PAH}$ values in the ordinate are expected to be substantially larger than the abscissa. However, the difference in both L$_{\\rm 3.3PAH}$ and EW$_{\\rm 3.3PAH}$ is relatively small, a factor of a few at most. We see no evidence that the spatially {\\it extended}, modestly obscured starbursts in host galaxies are substantially (more than an order of magnitude) more luminous than the modestly obscured {\\it nuclear} starbursts in ULIRGs. Even including extended starbursts in host galaxies, the L$_{\\rm 3.3 PAH}$/L$_{\\rm IR}$ ratios in ULIRGs are factors of 2.5 to more than 10 times smaller than those found in less infrared-luminous starbursts ($\\sim$10$^{-3}$; Mouri et al. 1990; Imanishi 2002). Taken at face value, the detected modestly obscured starbursts can account for only $<$10\\% to at most 40\\% of the infrared luminosities of the observed ULIRGs. Our AKARI IRC spectra reinforce the previous arguments, based on a small sample \\citep{soi00}, that nearby ULIRGs at $z <$ 0.3 are not energetically dominated by extended modestly obscured starburst activity in host galaxies. \\subsection{Buried AGNs with weak starbursts} The deficit in observed 3.3 $\\mu$m PAH luminosity relative to the infrared luminosity requires energy sources in addition to the modestly obscured (A$_{\\rm V}$ $<$ 15 mag) PAH-emitting starbursts. The first possibility is very highly obscured (A$_{\\rm V}$ $>>$ 15 mag) PAH-emitting starbursts, in which the 3.3 $\\mu$m PAH flux is severely attenuated, while the flux of longer wavelength infrared emission (8--1000 $\\mu$m) may not be highly attenuated. The second possibility is that an AGN exists that can produce large infrared dust emission luminosities with no PAH emission (see $\\S$1), decreasing the observed L$_{\\rm 3.3 PAH}$/L$_{\\rm IR}$ ratios. These two scenarios are difficult to differentiate based on the absolute PAH luminosities but can be distinguished by the {\\it equivalent width} of emission or absorption features. In a normal starburst galaxy, where HII regions, molecular gas, and photodissociation regions are spatially well mixed (Figure 1a), the equivalent width of the 3.3 $\\mu$m PAH-emission feature is insensitive to dust extinction \\citep{idm06}. If PAH-free AGN emission contributes significantly to the observed 2.5--5 $\\mu$m flux, then the EW$_{\\rm 3.3PAH}$ value should decrease. The EW$_{\\rm 3.3PAH}$ values in starbursts have an average value of EW$_{\\rm 3.3PAH}$ $\\sim$ 100 nm, with some scatter, but never become lower than 40 nm \\citep{moo86}. Thus, we adopt EW$_{\\rm 3.3PAH}$ $\\lesssim$ 40 nm as a strong signature of significant AGN contribution to an observed flux. Among ULIRGs classified optically as non-Seyferts in the IRAS 1 Jy sample, three LINER ULIRGs, IRAS 23129+2548, 08572+3915, and 17044+6720, and one HII-region ULIRG, IRAS 22088$-$1831, have EW$_{\\rm 3.3PAH}$ $<$ 20 nm (Table 3, column 5), more than a factor of 5 less than typical starburst galaxies. These ULIRGs are strong buried AGN candidates. The following non-Seyfert ULIRGs are also taken to contain luminous buried AGNs because the EW$_{\\rm 3.3PAH}$ values are $<$40 nm: the three LINER ULIRGs (IRAS 04074$-$2801, 11180+1623, and 21477+0502), two HII-region ULIRGs (IRAS 14202+2615 and 17028+5817E), and one optically unclassified ULIRG (IRAS 08591+5248). \\subsection{Buried AGNs with coexisting strong starbursts} Based on the EW$_{\\rm 3.3 PAH}$ values, we can easily detect buried AGNs with very weak starbursts. Even if strong starburst activity is present, {\\it weakly obscured} AGNs are detectable because weakly attenuated PAH-free continua from the AGNs can dilute the 3.3 $\\mu$m PAH emission considerably. However, detecting {\\it deeply buried} AGNs with coexisting surrounding strong starbursts is very difficult (Figure 1c). Even if the intrinsic luminosities of a buried AGN and surrounding less-obscured starbursts are similar, the AGN flux will be more highly attenuated by dust extinction than the starburst emission. When a buried AGN is obscured by {\\it ice-covered} dust grains, the AGN flux at $\\lambda_{\\rm rest}$ = 3.3 $\\mu$m is attenuated even more severely by the strong, broad 3.1 $\\mu$m absorption feature, making the EW$_{\\rm 3.3PAH}$ values in observed spectra apparently large. To determine whether a deeply buried AGN is present in addition to strong starbursts, we use the optical depths of dust absorption features found in the 2.5--5 $\\mu$m spectra. As described in $\\S$1 and in \\citet{idm06} in more detail, these values can be used to distinguish whether the energy sources are spatially well mixed with dust (a normal starburst), or are more centrally concentrated than the dust (a buried AGN). For a normal starburst with the mixed dust/source geometry in a ULIRG's core, $\\tau_{3.1}$ cannot exceed 0.3, while a buried AGN can produce $\\tau_{3.1} >$ 0.3 \\citep{im03,idm06}. Therefore, detection of $\\tau_{3.1} >$ 0.3 can be used to argue for the presence of a buried AGN with a centrally concentrated energy source geometry. Considering the uncertainty of $\\tau_{3.1}$ with $\\sim$0.1 ($\\S$4), we classify ULIRGs with $\\tau_{3.1}$ $>$ 0.4 as buried AGN candidates. Aside from the above ULIRGs with low EW$_{\\rm 3.3PAH}$, the following non-Seyfert ULIRGs in the IRAS 1 Jy sample are newly classified as buried AGNs: ten LINER ULIRGs (IRAS 05020$-$2941, 09463+8141, 11028+3130, 14121$-$0126, 16333+4630, 00482$-$2721, 09539+0857, 10494+4424, 16468+5200, 17028+5817W), four HII-region ULIRGs (IRAS 01199$-$2307, 01355$-$1814, 17068+4027, 11387+4116), and one optically unclassified ULIRG (IRAS 01494$-$1845). This large $\\tau_{3.1}$ method is sensitive to deeply buried AGNs but obviously misses weakly obscured AGNs, which are more easily detected with the above low EW$_{\\rm 3.3PAH}$ method. Hence, the large $\\tau_{3.1}$ and low EW$_{\\rm 3.3PAH}$ methods are used in complementary fashion to detect AGN signatures. We can also investigate the dust/source geometry from the $\\tau_{3.4}$ value. ULIRGs with $\\tau_{3.4} >$ 0.2 can be used to argue for the presence of buried AGNs \\citep{im03,idm06}. Among the optically non-Seyfert ULIRGs in the IRAS 1 Jy sample, only the LINER ULIRG IRAS 08572+3915 displays $\\tau_{3.4}$ $>$ 0.2. The LINER ULIRG of interest, UGC 5101, and the Seyfert 2 ULIRG, IRAS 19254$-$7245, also show $\\tau_{3.4}$ $>$ 0.2. However, all of these three ULIRGs have already been classified as having luminous AGNs based on their low EW$_{\\rm 3.3PAH}$ values. Imanishi et al. (2006a, 2007a) commented that exceptionally centrally concentrated starbursts (Figure 1d) and normal starbursts with mixed dust/source geometry obscured by foreground dust in edge-on host galaxies (Figure 1e) can also produce large $\\tau_{3.1}$ and $\\tau_{3.4}$ values, but argued that it is very unlikely for the bulk of ULIRGs with $\\tau_{3.1} >$ 0.3 and/or $\\tau_{3.4} >$ 0.2 to correspond to these non-AGN cases. For the remaining non-Seyfert ULIRGs with EW$_{\\rm 3.3PAH}$ $>$ 40 nm, $\\tau_{3.1}$ $\\lesssim$ 0.4, and $\\tau_{3.4}$ $\\lesssim$ 0.2, no obvious buried AGN signatures were observed in the 2.5--5 $\\mu$m spectra. Their spectra can be explained by either of the following scenarios: normal starbursts with mixed dust/source geometry are energetically dominant, or AGNs are present, but the AGN emission is so highly attenuated that its contribution to the observed 2.5--5 $\\mu$m flux is not significant. We have no way of distinguishing between these two scenarios. However, some examples are known (e.g., NGC 4418) in which buried AGN signatures are found only at wavelengths longer than 5 $\\mu$m \\citep{ima04,dud97,spo01}. Thus, the detected buried AGN fraction in the 2.5--5 $\\mu$m AKARI IRC spectra is only a lower limit. \\subsection{Dust extinction and intrinsic AGN luminosities} In a buried AGN with centrally concentrated energy source geometry, dust at a temperature of 1000 K, which is close to the innermost dust sublimation radius, produces continuum emission with a peak at $\\lambda \\sim$ 3 $\\mu$m, assuming approximately blackbody emission. Since a foreground screen dust distribution model is applicable to a buried AGN (Imanishi et al. 2006a, 2007a), the $\\tau_{3.1}$ (ice-covered dust) and $\\tau_{3.4}$ (bare dust) values reflect the dust column density toward the 3 $\\mu$m continuum-emitting region, which is almost equal to the column density toward the buried AGN itself (Imanishi et al. 2006a, 2007a). The 3.1 $\\mu$m absorption feature is detectable if ice-covered dust grains are present in front of the 3--4 $\\mu$m continuum emitting energy source. Such ice-covered dust grains are usually found deep inside molecular gas, where ambient UV radiation is sufficiently shielded \\citep{whi88,tan90,smi93,mur00}. The 3.4 $\\mu$m absorption feature should be detected if the 3--4 $\\mu$m continuum emitting energy source is obscured by bare carbonaceous dust grains \\citep{pen94,ima96,raw03}, but is undetected if the absorbing dust is ice-covered \\citep{men01}. Since absorbing dust consists of both bare and ice-covered dust grains, ULIRGs with obscured energy sources should show both features, and the true dust column density is derivable from a proper combination of $\\tau_{3.1}$ and $\\tau_{3.4}$. However, despite the detection of the 3.1 $\\mu$m ice absorption features in many ULIRGs, only two ULIRGs in the IRAS 1 Jy sample, IRAS 08572+3915 and 17044+6720, show clear 3.4 $\\mu$m carbonaceous dust absorption features. Even including the two sources of interest, UGC 5101 and IRAS 19254$-$7245, only four ULIRGs display clearly detectable 3.4 $\\mu$m absorption features. The difference in the detection rate largely comes from the intrinsically smaller oscillator strength of the 3.4 $\\mu$m carbonaceous dust absorption feature ($\\tau_{3.4}$/A$_{\\rm V}$ = 0.004--0.007; Pendleton et al. 1994) compared to the 3.1 $\\mu$m ice absorption feature ($\\tau_{3.1}$/A$_{\\rm V}$ = 0.06; Tanaka et al. 1990; Smith et al. 1993; Murakawa et al. 2000). Even if a modestly large amount of bare carbonaceous dust grains is present in front of the continuum-emitting energy source, the $\\tau_{3.4}$ value is small, making the detection of the 3.4 $\\mu$m dust absorption feature difficult. Additionally, the 3.3 $\\mu$m PAH emission feature is often accompanied by a sub-peak at 3.4 $\\mu$m \\citep{tok91,imd00}, and this sub-peak may dilute the 3.4 $\\mu$m dust absorption feature at the same wavelength. In fact, all the four ULIRGs with detectable 3.4 $\\mu$m absorption features are limited to relatively weak PAH emitters (EW$_{\\rm 3.3PAH}$ $<$ 35 nm). Finally, when the spectrally broad 3.1 $\\mu$m absorption feature is strong, the absorption feature extends to the longer wavelength side of the 3.3 $\\mu$m PAH emission feature, making it difficult to distinguish the origin of the apparent flux depression at $\\lambda_{\\rm rest}$ $\\sim$ 3.4 $\\mu$m. Among the four sources with detectable 3.4 $\\mu$m absorption features, IRAS 08572+3915, 17044+6720, and 19254$-$7245 indeed display only weak or undetectable 3.1 $\\mu$m absorption features. The remaining source, UGC 5101, shows large $\\tau_{3.1}$, but we can recognize the 3.4 $\\mu$m absorption feature, primarily because UGC 5101 is one of the brightest sources and the signal to noise ratios are among the highest. Detection of the 3.4 $\\mu$m absorption features in the remaining ULIRGs with large EW$_{\\rm 3.3PAH}$, large $\\tau_{3.1}$, and limited signal-to-noise ratios in the continuum is basically difficult. Therefore, while the total dust column density can be estimated in a reasonably reliable way for ULIRGs with both detectable 3.1 $\\mu$m and 3.4 $\\mu$m absorption features, the estimated dust column densities are only lower limits, and should be much smaller than the actual values for ULIRGs with only detectable 3.1 $\\mu$m absorption features. The estimated dust column densities are summarized in column 4 of Table 4. In a buried AGN, the surrounding dust has a strong temperature gradient in that inner dust, close to the central energy source, has higher temperature than outer dust. Luminosity is transferred at each temperature, and the intrinsic luminosity of inner hot dust emission at 3--4 $\\mu$m ($\\nu$F$_\\nu$) should be comparable to that of outer cool dust emission at 60 $\\mu$m, the wavelength which dominates the observed infrared emission of ULIRGs \\citep{san88a}. Thus, if the intrinsic AGN's 3--4 $\\mu$m luminosity ($\\nu$F$_\\nu$) is comparable to the observed infrared luminosities of ULIRGs, then we can argue that the buried AGN is energetically important. For ULIRGs with low EW$_{\\rm 3.3PAH}$ ($<$40 nm), we can roughly extract the AGN's PAH-free continuum at 3--4 $\\mu$m, based on the assumption that starburst activity intrinsically shows EW$_{\\rm 3.3PAH}$ $\\sim$ 100 nm. That is, for ULIRGs with EW$_{\\rm 3.3PAH}$ = 30 nm (30\\% of the typical starburst value), we consider that 70\\% of the 3--4 $\\mu$m continuum comes from AGN's PAH-free continuum emission. Thus, we can estimate the {\\it observed } 3--4 $\\mu$m flux of AGN emission. Next, for ULIRGs with both detectable 3.1 $\\mu$m and 3.4 $\\mu$m absorption features (IRAS 08572+3915, UGC 5101, and IRAS 19254$-$7245), and for IRAS 17044+6720, which displays only the 3.4 $\\mu$m absorption feature, we can estimate dust column density, or {\\it dust extinction}, toward the 3--4 $\\mu$m continuum emitting regions based on $\\tau_{3.1}$ and $\\tau_{3.4}$. When we combine the {\\it observed} AGN flux at 3--4 $\\mu$m and {\\it dust extinction} toward the 3--4 $\\mu$m continuum emitting regions, we can derive the dust-extinction-corrected {\\it intrinsic} AGN flux, and thus the intrinsic AGN luminosity. If we adopt A$_{\\rm 3-4 \\mu m}$/A$_{\\rm V}$ $\\sim$ 0.058 \\citep{rie85}, then we find $\\tau_{3.1}$/A$_{\\rm 3-4 \\mu m}$ $\\sim$ 1 and $\\tau_{3.4}$/A$_{\\rm 3-4 \\mu m}$ $\\sim$ 0.069--0.12 for the Galactic interstellar medium. Assuming this relationship, we estimate the intrinsic AGN luminosities at 3--4 $\\mu$m ($\\nu$F$_{\\nu}$) to be L $\\gtrsim$ 10$^{12}$L$_{\\odot}$ in all the 3.4$\\mu$m-absorption-detected ULIRGs. The putative AGN activity is therefore energetically sufficient to quantitatively account for the bulk of the infrared luminosities of these ULIRGs (L$_{\\rm IR}$ $\\sim$ 10$^{12}$L$_{\\odot}$). \\subsection{Dependence of the buried AGN fraction on optical spectral type: LINER vs. HII-region} In total, AKARI IRC 2.5--5 $\\mu$m spectra of 19 LINER, 16 HII-region, and 5 optically unclassified ULIRG's nuclei in the {\\it IRAS} 1 Jy sample were obtained. The low EW$_{\\rm 3.3PAH}$ method suggests six LINER and three HII-region ULIRG's nuclei contain luminous buried AGNs ($\\S$5.2). In addition to these ULIRGs, the large $\\tau_{3.1}$ method classifies ten LINER and four HII-regions ULIRG's nuclei as sources with deeply obscured buried AGNs ($\\S$5.3). In total, the detected buried AGN fraction is 16/19 (84\\%) for LINER ULIRGs, and 7/16 (44\\%) for HII-region ULIRGs (Table 4, columns 5 and 6). Since the selection of the observed ULIRGs is based solely on the target's visibility from the AKARI satellite and should be unbiased in terms of their energy sources ($\\S$2), we argue that the detected buried AGN fraction is higher in LINER ULIRGs than in HII-region ULIRGs. The same result was found from ground-based $L$-band spectroscopy and Spitzer 5--35 $\\mu$m spectroscopy of ULIRGs at $z <$ 0.15 (Imanishi et al. 2006a, 2007a), and also from a VLA radio observational search for compact radio core emission (another good AGN indicator) \\citep{nag03}. Therefore, we confirm that a larger fraction of LINER ULIRGs possess luminous buried AGNs than HII-region ULIRGs with a probability of $\\sim$99\\%. The higher buried AGN fraction in optically LINER ULIRGs can be explained qualitatively by a dustier starburst scenario \\citep{ima07a}. For starburst/buried AGN composite ULIRGs (Figure 1c), the optical LINER or HII-region classifications are likely largely affected by the properties of the modestly obscured starbursts at the exteriors of the buried AGN, rather than by buried AGN-related emission, as optical observations can probe only the surfaces of dusty objects. In a dusty starburst, shock-related emission can be relatively important in the optical compared to the emission from the HII-regions themselves, resulting in optical LINER classification. When a luminous AGN is placed at the center of a {\\it less dusty} starburst classified optically as an HII-region, the AGN emission is more easily detectable in the optical, making such an object an optical Seyfert. In contrast, when a luminous AGN is placed at the center of a {\\it dusty} starburst classified optically as a LINER, the AGN emission is more elusive in the optical, so that such an object is classified as an optical non-Seyfert. Hence, this scenario can explain the observed higher fraction of optically elusive {\\it buried} AGNs in optically LINER ULIRGs compared to HII-region ULIRGs. In fact, the emission probed in the optical was found to be dustier in LINER ULIRGs than in HII-region ULIRGs (Veilleux et al. 1995, 1999). \\subsection{The buried AGN fraction as a function of infrared luminosity} Due to the inclusion of ULIRGs at $z >$ 0.15, we have now a large number of ULIRGs with L$_{\\rm IR}$ $\\gtrsim$ 10$^{12.3}$L$_{\\odot}$. Specifically, only 1 of 13 observed non-Seyfert ULIRGs at $z <$ 0.15 has L$_{\\rm IR}$ $\\gtrsim$ 10$^{12.3}$L$_{\\odot}$, while 16 of 26 observed non-Seyfert ULIRGs at $z >$ 0.15 have L$_{\\rm IR}$ $\\gtrsim$ 10$^{12.3}$L$_{\\odot}$ (see Table 1). We can thus investigate the buried AGN fraction, separating ULIRGs into two categories: those with L$_{\\rm IR}$ $<$ 10$^{12.3}$L$_{\\odot}$ and those with L$_{\\rm IR}$ $\\gtrsim$ 10$^{12.3}$L$_{\\odot}$. Among the observed non-Seyfert ULIRGs in the IRAS 1 Jy sample, 22 sources have L$_{\\rm IR}$ $<$ 10$^{12.3}$L$_{\\odot}$ and the remaining 17 sources have L$_{\\rm IR}$ $\\gtrsim$ 10$^{12.3}$L$_{\\odot}$. Based on the buried AGN signatures in Table 4, the fraction of ULIRGs with detectable buried AGN signatures is 12/22 (= 55\\%) for non-Seyfert ULIRGs with L$_{\\rm IR}$ $<$ 10$^{12.3}$L$_{\\odot}$ and 12/17 (= 71\\%) for non-Seyfert ULIRGs with L$_{\\rm IR}$ $\\gtrsim$ 10$^{12.3}$L$_{\\odot}$. Although the total sample size is not large, we find a higher buried AGN fraction with increasing ULIRG infrared luminosity. \\citet{idm06} investigated, based on ground-based $L$-band (2.8--4.1 $\\mu$m) spectra, the buried AGN fraction in a larger number of ULIRGs at $z <$ 0.15 in the IRAS 1 Jy sample than in this paper. Several ULIRGs at $z <$ 0.15 studied in this paper are also included in this ground-based study. When we divide the observed ULIRGs by \\citet{idm06} into those with L$_{\\rm IR}$ $<$ 10$^{12.3}$L$_{\\odot}$ and $\\gtrsim$ 10$^{12.3}$L$_{\\odot}$, the detected buried AGN fraction is 16/29 (= 55\\%) in non-Seyfert ULIRGs with L$_{\\rm IR}$ $<$ 10$^{12.3}$L$_{\\odot}$, and 6/9 (= 67\\%) in non-Seyfert ULIRGs with L$_{\\rm IR}$ $\\gtrsim$ 10$^{12.3}$L$_{\\odot}$. The detected buried AGN fractions in ground-based and AKARI IRC spectra are comparable for non-Seyfert ULIRGs with both L$_{\\rm IR}$ $<$ 10$^{12.3}$L$_{\\odot}$ ($\\sim$55\\%) and $\\gtrsim$ 10$^{12.3}$L$_{\\odot}$ ($\\sim$70\\%). We therefore argue that the detected buried AGN fraction increases with ULIRG infrared luminosity. This trend parallels the higher detection rate of optical Seyfert signatures in ULIRGs with higher infrared luminosities \\citep{vei99}. The fraction of both optical Seyferts and luminous buried AGNs is significantly smaller in galaxies with L$_{\\rm IR}$ $<$ 10$^{12}$L$_{\\odot}$ than ULIRGs \\citep{vei99,soi01}. Hence, we conclude that AGN activity becomes more important as the infrared luminosities of galaxies increase. Recently, the so-called galaxy downsizing phenomena have been found, where galaxies with currently larger stellar masses have finished their major star-formation in earlier cosmic age \\citep{cow96,bun05}. AGN feedbacks are proposed to be responsible for the galaxy downsizing phenomena \\citep{gra04,bow06,cro06}. Namely, in galaxies with currently larger stellar masses, AGN feedbacks have been stronger in the past, and gas has been expelled in a shorter time scale. Buried AGNs can have particularly strong feedbacks, because the AGNs are surrounded by a large amount of nuclear gas and dust. If we reasonably assume that galaxies with currently larger stellar masses have previously been more infrared luminous, then the detected higher buried AGN fraction in more infrared luminous galaxies may support the AGN feedback scenario as the origin of the galaxy downsizing phenomena \\footnote{ To form more stars, more star-formation should have been occurred in the past, producing stronger star-formation related infrared emission in the past. If AGN's energetic contribution is negligible in LIRGs with L$_{\\rm IR}$ $<$ 10$^{12}$L$_{\\odot}$, but important, say $\\sim$50\\%, in ULIRGs with L$_{\\rm IR}$ $>$ 10$^{12}$L$_{\\odot}$, then AGN feedbacks can be stronger in ULIRGs, and yet higher infrared luminosity ($\\sim$5 $\\times$ 10$^{11}$L$_{\\odot}$) can come from star-forming activity in ULIRGs, producing more stellar masses in ULIRGs than in LIRGs. }. \\subsection{Comparison with Seyfert 2 ULIRGs} We compare the 2.5--5 $\\mu$m spectral properties of non-Seyfert ULIRGs showing buried AGN signatures with those of Seyfert 2 ULIRGs (i.e., known obscured-AGN-possessing ULIRGs). The apparent main difference between them is that the ionizing radiation from the putative buried AGNs in non-Seyfert ULIRGs is obscured by the surrounding dust along virtually all lines-of-sight; in contrast, dust around the AGNs in Seyfert 2 ULIRGs is distributed in a ``torus'', and ionizing radiation from the AGNs can escape along the torus axis, allowing for optical Seyfert signature detection (Figure 5). The four ULIRGs classified optically as Seyfert 2s in the IRAS 1 Jy sample show no clear absorption features at 2.5--5 $\\mu$m. \\citet{idm06} also found in ground-based $L$-band spectra that the fraction of optically classified Seyfert 2 ULIRGs showing strong dust absorption features is substantially smaller than buried AGNs in optically non-Seyfert ULIRGs. We thus argue that the line-of-sight dust column density toward the AGNs is lower in Seyfert 2 ULIRGs than buried AGNs in non-Seyfert ULIRGs. Thus, buried AGNs and Seyfert 2 AGNs (obscured by torus-shaped dust) differ not only in dust geometry, but also in dust column density along our line-of-sight; specifically, the dust columns toward the buried AGNs are much higher \\citep{idm06}. Since the dust covering factor around buried AGNs (almost all directions) is also larger than Seyfert 2 AGNs (torus-shaped), the total amount of nuclear dust must be larger in the former (Figure 5). Since the gas and dust in an AGN have angular momentum with respect to the central supermassive black hole, an axisymmetric spatial distribution is more natural than a spherical geometry. In this case, the column densities can be high in certain directions but low in others (Figure 5). For a fixed angular momentum, the dust column density ratios between the highest and lowest column directions are similar among different galaxies. If the total amount of nuclear dust is modest, the direction of the lowest dust column density can be transparent to the AGN's ionizing radiation, making Seyfert signatures detectable in the optical spectra. As the total amount of nuclear dust increases, even the direction of the lowest dust column density can be opaque to the AGN's ionizing radiation, making such galaxies optically elusive buried AGNs. Thus, all of the observed spectral properties of buried AGNs and Seyfert-type AGNs are explicable given a larger amount of nuclear dust in the former. Since ULIRGs contain a large amount of nuclear gas and dust \\citep{sam96}, the high buried AGN fraction is inevitable. Understanding optically elusive buried AGNs is therefore essential if we are to unveil the true nature of the ULIRG population. \\subsection{Dependence on far-infrared colors} Based on the {\\it IRAS} 25 $\\mu$m to 60 $\\mu$m flux ratio ($f_{\\rm 25}$/$f_{\\rm 60}$), ULIRGs are divided into cool ($<$ 0.2) and warm ($>$ 0.2) sources \\citep{san88b}. Many of the non-Seyfert ULIRGs with detectable buried AGN signatures show cool far-infrared colors (Table 1). Although AGNs classified optically as Seyferts usually show warm far-infrared colors \\citep{deg87,kee05}, cool far-infrared colors of buried AGNs are the natural consequence of a large amount of nuclear dust, where contributions from the outer, cooler dust components to the infrared radiation become more important than for optical Seyferts. Figure 6(a) compares {\\it IRAS} 25 $\\mu$m to 60 $\\mu$m flux ratios (i.e., far-infrared color) with the observed 3.3 $\\mu$m PAH to infrared luminosity ratios (L$_{\\rm 3.3PAH}$/L$_{\\rm IR}$). Seyfert ULIRGs tend to appear in the warmer far-infrared color range than non-Seyfert ULIRGs, as expected from the decreased amount of nuclear dust in the former ($\\S$5.7). No systematic difference in the L$_{\\rm 3.3PAH}$/L$_{\\rm IR}$ ratios between non-Seyfert and Seyfert ULIRGs exists. Figure 6(b) compares the far-infrared colors and EW$_{\\rm 3.3PAH}$ for non-Seyfert ULIRGs. Buried AGNs appear in both the warm and cool ranges. Although it is sometimes argued that ULIRGs with cool far-infrared colors must be starburst-dominated, simply because Seyfert-type AGNs show warm far-infrared colors (e.g., Downes \\& Solomon 1998), we do not confirm this to be true for the heavily buried AGNs." }, "0808/0808.0539_arXiv.txt": { "abstract": "Using a large galaxy group catalogue constructed from the Sloan Digital Sky Survey Data Release 4 (SDSS DR4) with an adaptive halo-based group finder, we investigate the luminosity and stellar mass functions for different populations of galaxies (central versus satellite; red versus blue; and galaxies in groups of different masses) and for groups themselves. The conditional stellar mass function (CSMF), which describes the stellar distribution of galaxies in halos of a given mass for central and satellite galaxies can be well modeled with a log-normal distribution and a modified Schechter form, respectively. On average, there are about 3 times as many central galaxies as satellites. Among the satellite population, there are in general more red galaxies than blue ones. For the central population, the luminosity function is dominated by red galaxies at the massive end, and by blue galaxies at the low mass end. At the very low-mass end ($M_\\ast \\la 10^9 h^{-2}\\Msun$), however, there is a marked increase in the number of red centrals. We speculate that these galaxies are located close to large halos so that their star formation is truncated by the large-scale environments. The stellar-mass function of galaxy groups is well described by a double power law, with a characteristic stellar mass at $\\sim 4\\times 10^{10}h^{-2}\\Msun$. Finally, we use the observed stellar mass function of central galaxies to constrain the stellar mass - halo mass relation for low mass halos, and obtain $M_{\\ast, c}\\propto M_h^{4.9}$ for $M_h \\ll 10^{11} \\msunh$. ", "introduction": "In recent years, great progress has been made in our understanding about how galaxies form and evolve in dark matter halos owing to the development of halo models and the related halo occupation models. The halo occupation distribution (hereafter HOD), $P(N \\vert M_h)$, which gives the probability of finding $N$ galaxies (with some specified properties) in a halo of mass $M_h$, has been extensively used to study the galaxy distribution in dark matter halos and galaxy clustering on large scales (e.g. Jing, Mo \\& B\\\"orner 1998; Peacock \\& Smith 2000; Seljak 2000; Scoccimarro \\etal 2001; Jing, B\\\"orner \\& Suto 2002; Berlind \\& Weinberg 2002; Bullock, Wechsler \\& Somerville 2002; Scranton 2002; Zehavi \\etal 2004, 2005; Zheng \\etal 2005; Tinker \\etal 2005; Skibba et al. 2007; Brown et al. 2008). The conditional luminosity function (hereafter CLF), $\\Phi(L \\vert M_h) {\\rm d}L$, which refines the HOD statistic by considering the average number of galaxies with luminosity $L \\pm {\\rm d}L/2$ that reside in a halo of mass $M_h$, has also been investigated extensively (Yang, Mo \\& van den Bosch 2003; van den Bosch, Yang \\& Mo 2003; Vale \\& Ostriker 2004, 2008; Cooray 2006; van den Bosch et al. 2007) and has been applied to various redshift surveys, such as the 2-degree Field Galaxy Redshift Survey (2dFGRS), the Sloan Digital Sky Survey (SDSS) and DEEP2 (e.g. Yan, Madgwick \\& White 2003; Yang \\etal 2004; Mo et al. 2004; Wang \\etal 2004; Zehavi \\etal 2005; Yan, White \\& Coil 2004). These investigations demonstrate that the HOD/CLF statistics are very powerful tools to establish and describe the connection between galaxies and dark matter halos, providing important constraints on various physical processes that govern the formation and evolution of galaxies, such as gravitational instability, gas cooling, star formation, merging, tidal stripping and heating, and a variety of feedback processes, and how their efficiencies scale with halo mass. Furthermore, they also indicate that the galaxy/dark halo connection can provide important constraints on cosmology (e.g.,van den Bosch, Mo \\& Yang 2003; Zheng \\& Weinberg 2007). However, as pointed out in Yang et al. (2005c; hereafter Y05c), a shortcoming of the HOD/CLF models is that the results are not completely model independent. Typically, assumptions have to be made regarding the functional form of either $P(N \\vert M_h)$ or $\\Phi(L \\vert M_h)$. Moreover, in all HOD/CLF studies to date, the occupation distributions have been determined in an indirect way: the free parameters of the assumed functional form are constrained using {\\it statistical} data on the abundance and clustering properties of the galaxy population. An alternative method that can directly probe the galaxy - dark halo connection (e.g. HOD/CLF models) is to use galaxy groups as a representation of dark matter halos and to study how the galaxy population changes with the properties of the groups (e.g., Y05c; Zandivarez et al. 2006; Robotham et al. 2006; Hansen \\etal 2007; Yang et al. 2008). For such a purpose, one has to properly find the galaxy groups that are closely connected to the dark matter halos. In recent studies, Yang et al. (2005a; 2007) developed an adaptive halo-based group finder that has such features \\footnote{In this paper, we refer to systems of galaxies as groups regardless of their richness, including isolated galaxies (i.e., systems with a single member) and rich clusters of galaxies.}. This group finder has been applied to the 2dFGRS (Yang et al. 2005a) and to the SDSS (Weinmann et al. 2006a; Yang et al. 2007). Detailed tests with mock galaxy catalogues have shown that this group finder is very successful in associating galaxies according to their common dark matter halos. In particular, the group finder performs reliably not only for rich systems, but also for poor systems, including isolated central galaxies in low mass halos. This makes it possible to study the galaxy-halo connection for systems covering a large dynamic range in masses. With a well-defined galaxy group catalogue, one can then not only study the properties of galaxies in different groups (e.g. Y05c; Yang \\etal 2005d; Collister \\& Lahav 2005; van den Bosch \\etal 2005; Robotham \\etal 2006; Zandivarez \\etal 2006; Weinmann \\etal 2006a,b; van den Bosch \\etal 2008; McIntosh \\etal 2007; Yang et al. 2008), but also probe how dark matter halos trace the large-scale structure of the Universe (e.g. Yang \\etal 2005b, 2006; Coil \\etal 2006; Berlind \\etal 2007; Wang et al. 2008a). Recently, this group finder has been applied to the Sloan Digital Sky Survey Data Release 4 (SDSS DR4), and the group catalogues constructed are described in detail in Yang \\etal (2007; Paper I hereafter). In these catalogues various observational selection effects are taken into account, and each of the groups is assigned a reliable halo mass. The group catalogues including the membership of the groups are available at these links \\footnote{http://gax.shao.ac.cn/data/Group.html} \\footnote{http://www.astro.umass.edu/$^\\sim$xhyang/Group.html}. In Yang et al. (2008; Paper II hereafter) we have used these group catalogues to obtain various halo occupation statistics and to measure the CLFs for different populations of galaxies. In this paper, the third in the series, we will focus on the conditional stellar mass functions (CSMFs) for different populations of galaxies. In addition, we will also examine the general luminosity and stellar mass functions for different populations of galaxies and for groups themselves. Finally, we will demonstrate how to use the observed luminosity and stellar mass functions for central galaxies to constrain the HOD in small halos. This paper is organized as follows: In Section~\\ref{sec_data} we describe the data (galaxy and group catalogues) used in this paper. Section~\\ref{sec_CSMFs} presents our measurement of the CSMFs for all, red and blue galaxies. Sections~\\ref{sec_LF_gax} and ~\\ref{sec_LF_grp} present our measurement of the luminosity and stellar mass functions for galaxies and groups, respectively. In Section \\ref{sec_small}, we probe the properties of the central galaxies that can be formed in those small halos. Finally, we summarize our results in Section~\\ref{sec_summary}. Throughout this paper, we use a $\\Lambda$CDM `concordance' cosmology whose parameters are consistent with the three-year data release of the WMAP mission: $\\Omega_m = 0.238$, $\\Omega_{\\Lambda}=0.762$, $n_s=0.951$, $h=0.73$ and $\\sigma_8=0.75$ (Spergel et al. 2007). If not quoted, the units of luminosity, stellar and halo masses are in terms of $h^{-2}\\Lsun$, $h^{-2}\\Msun$ and $h^{-1}\\Msun$, respectively. Finally, unless noted differently, the luminosity functions and stellar mass functions are presented in units of $h^3{\\rm Mpc}^{-3} {\\rm d} \\log L$ and $h^3{\\rm Mpc}^{-3} {\\rm d} \\log M_{\\ast}$, respectively, where $\\log$ is the 10 based logarithm. ", "conclusions": "\\label{sec_summary} In this paper, we have derived the luminosity and stellar mass functions for different populations of galaxies (central versus satellite; red versus blue; and galaxies in halos of different masses), and for groups themselves, using a large galaxy group catalogue constructed from the SDSS Data Release 4 (DR4). Our main results can be summarized as follows: \\begin{enumerate} \\item For central galaxies, the conditional stellar mass function (CSMF), which describes the stellar mass distribution of galaxies in halos of a given mass can be well described by a log-normal distribution, with a width $\\sigma_{\\rm log M_\\ast}\\sim 0.17$, quite independent of the host halo mass. The median central stellar mass increases rapidly with halo mass, $M_\\ast\\propto M_h^{4.9}$, for halos with masses $M_h\\ll 10^{11}\\msunh$, but only slowly, $M_\\ast\\propto M_h^{0.3}$, for halos with $M_h\\gg 10^{13}\\msunh$. \\item For satellite galaxies, the conditional stellar mass function in halos of different masses can be described reasonably well by a modified Schechter form than breaks away faster than the Schechter function at the massive end. The faint end slope appears to be steeper for more massive halos. On average, there are about 3 times as many central galaxies as satellites. \\item When stellar mass functions are measured separately for galaxies of different colors, we find that the central population is dominated by red galaxies at the massive end, and by blue galaxies at the low-mass end. Among the satellite population, there are in general more red galaxies than blue ones. At the very low-mass end ($M_\\ast \\la 10^9 h^{-2}\\Msun$), there is a marked increase in the number of red centrals. We speculate that these galaxies are located close to large halos so that their star formation has been affected by their environments. \\item The stellar-mass function of galaxy groups, which describes the number density of galaxy groups as a function of the total stellar mass of group member galaxies, is well described by a double power law, with a characteristic stellar mass at $\\sim 4\\times 10^{10}h^{-2}\\Msun$. This form is very different from that of the halo mass function, indicating that the efficiencies of star formation in halos of different masses are very different. \\item The stellar mass function for the central galaxies can be used to provide stringent constraint on the mean $M_{\\ast,c}$ - $M_h$ relation for low-mass halos. \\end{enumerate} We anticipate that a comparison of these results with predictions of numerical simulations and/or semi-analytical models will provide stringent constraints on how galaxies form and evolve in dark matter halos." }, "0808/0808.0822_arXiv.txt": { "abstract": "{ Though pick-up ions (PUIs) are a well known phenomenon in the inner heliosphere, their phase-space distribution nevertheless is a theoretically unsettled problem. Especially the question of how pick-up ions form their suprathermal tails, extending to far above their injection energies, still now is unsatistactorily answered. Though Fermi-2 velocity diffusion theories have revealed that such tails are populated, they nevertheless show that resulting population densities are much less than seen in observations showing power-laws with a velocity index of ``-5''. We first investigate here, whether or not observationally suggested power-laws can be the result of a quasi-equilibrium state between suprathermal ions and magnetohydrodynamic turbulences in energy exchange with eachother. We demonstrate that such an equilibrium cannot be established, since it would require too high pick-up ion pressures enforcing a shock-free deceleration of the solar wind. We furthermore show that Fermi-2 type energy diffusion in the outer heliosphere is too inefficient to determine the shape of the distribution function there. As we can show, however, power-laws beyond the injection threshold can be established, if the injection takes place at higher energies of the order of 100 keV. As we demonstrate here, such an injection is connected with modulated anomalous cosmic ray (ACR) particles at the lower end of their spectrum when they again start being convected outwards with the solar wind. Therefore, we refer to these particles as ACR-PUIs. In our quantitative calculation of the pick-up ion spectrum resulting under such conditions we in fact find again power-laws, however with a velocity power index of ``-4'' and fairly distance-independent spectral intensities. As it seems these facts are observationally well supported by VOYAGER measurements in the lowest energy channels. ", "introduction": "Suprathermal ions, picked-up by the supersonic solar wind flow as ionized neutral atoms, have become known as pick-up ions (PUIs) and are produced all over the inner heliosphere with a typical upwind-downwind asymmetry with respect to the inflow direction of the neutral ISM inflow vector \\citep{rucinski93,fahr99}. In the case of PUI protons, their production is due to photoionization and charge exchange of interstellar H-atoms \\citep[see][]{rucinski91,fahr99,rucinski03,bzowski07}. Their spatial distribution seems well understood, while the PUI phase-space transport is a much less settled subject. Especially it exists an ongoing debate of how efficiently pick-up ions just after the pick-up process are accelerated to higher energies due to nonlinear wave-particle interactions \\citep[see e.g.][]{isenberg87,bogdan91,fichtner96,fichtner01,chalov96,chalov98,chalov04} and whether at all energy diffusion plays a relevant role in this transport. Some hint is given by the solar wind proton temperature behavior with distance. The observed non-adiabatic temperature behavior namely proves that a specific solar wind proton heating must operate in the outer heliosphere which can only be due to energy absorption from pick-up ion generated turbulence, since convected turbulence amplitudes quickly die out with distance \\citep[see][]{smith01,chashei02}. Freshly injected PUIs represent keV-energetic protons in the supersonic solar wind frame and may be called here: ``primary pick-up ions'' (or: PUIs$^{\\ast }$). The velocity distribution of these newly produced PUIs$^{\\ast }$ is toroidal and unstable \\citep[see][]{winske84,winske85,lee87,fahr88}. With the free energy of this unstable distribution PUIs$^{\\ast }$ drive Alfv\\'{e}nic wave power. The latter enforces pitchangle isotropization of the initial velocity distribution \\citep[see][]{chalov98,chalov99}. Due to wave-wave coupling, the wave energy generated by PUIs$^{\\ast }$ at the injection wavelength $\\lambda _{i}=U_{s}/\\Omega _{p} $ is diffusively transported in wavevector space both to smaller wavelengths where it can be absorbed by solar wind protons and to larger wavelengths where it is reabsorbed by all PUIs. This effect is seen as the main reason of solar wind proton heating occuring in the outer heliosphere \\citep{smith01,fahr02,chashei03,stawicki04}. Only a small fraction of about 5 percent of the PUI-generated wave energy reappears in the observed proton temperatures. Freshly injected PUIs excite turbulences that can organize a power-law distribution. From this distribution, both the solar wind ions and the PUIs themselves can absorb energy as shown by \\citet{chashei03}. Also the approach by \\citet{isenberg03} where energy diffusion of pick-up ions is not taken into account shows that only a low degree (2-5 percent) of the pick-up ion driven wave energy is absorbed by solar wind protons in form of thermal energy. This raises the question where the major portion of the wave energy produced during the primary pick-up process goes to. To clarify the energy redistributions, kinetic and spectral details of the relevant processes have to be investigated. A detailed numerical study of the PUI velocity distribution and the spectral Alfv\\'{e}nic/Magnetosonic wave power evolution has meanwhile been carried out \\citep{chalov04,chalov06a} and presents a simultaneous solution of a coupled system of equations consistently describing the isotropic velocity distribution function of PUIs and the spectral wave power intensity. As one can see from this study, the largest portion of the self-generated wave energy is reabsorbed by PUIs themselves as a result of the cyclotron resonant interaction and leads to PUI-acceleration. It could perhaps be hoped that this energization of pick-up protons due to Fermi-2 stochastic acceleration processes eventually leads to the ubiquitous power-law PUI-tails pointed out by \\citet{fisk06,fisk07}. To the opposite, however, as reflected in the results presented by \\citet{chalov04,chalov06a} it is evident that this is not the case: even at larger distances close to the termination shock (100 AU) the PUI distributions show a rapid cut-off at energies higher than the injection energy. The question thus is raised here why power-laws have been seen at all. An explanation that we are favoring here is a new injection source to the PUI regime from high energies connected with modulated anomalous cosmic ray particles. These protons are primary ACR particles that occur with a spectrum down to the typical energy of the usually assumed PUIs. At this part of the spectrum, both particle species cannot be distinguished. Therefore, we refer to them as ACR-PUIs. In Sect.~\\ref{sect2} we investigate the physical possibility of power-law ions in the outer heliosphere as they are recently proposed by several authors and we find that they cannot occur with a power-index of $-5$. As we show in Sect.~\\ref{sect3}, the proposed processes are not effective enough to produce the desired ion tails which means that another mechanism has to lead to the observed spectrum. In Sect.~\\ref{sect4}, we show how a high-energy source can be derived by taking a modulated ACR spectrum upstream of the solar wind termination shock. This injection mechanism is discussed in Sect.~\\ref{sect5} where we show that these high-energy ions can lead to power-law ion tails, however with a power-index of $-4$. The results are discussed and compared with observations in Sect.~\\ref{sect6} ", "conclusions": "" }, "0808/0808.2016_arXiv.txt": { "abstract": "We review the present status of three-flavour neutrino oscillations, taking into account the latest available neutrino oscillation data presented at the {\\em Neutrino 2008\\/} Conference. This includes the data released this summer by the MINOS collaboration, the data of the neutral current counter phase of the SNO solar neutrino experiment, as well as the latest KamLAND and Borexino data. We give the updated determinations of the leading 'solar' and 'atmospheric' oscillation parameters. We find from global data that the mixing angle $\\theta_{13}$ is consistent with zero within $0.9\\sigma$ and we derive an upper bound of $\\sin^2\\theta_{13} \\leq 0.035 \\, (0.056)$ at 90\\%~CL (3$\\sigma$). \\vskip 1cm \\noindent Keywords: Neutrino mass and mixing; solar and atmospheric neutrinos; reactor and accelerator neutrinos ", "introduction": "\\label{sec:introduction} Thanks to the synergy amongst a variety of experiments involving solar and atmospheric neutrinos, as well as man-made neutrinos at nuclear power plants and accelerators~\\cite{exp-talks-nu08} neutrino physics has undergone a revolution over the last decade or so. The long-sought-for phenomenon of neutrino oscillations has been finally established, demonstrating that neutrino flavor states $(\\nu_e,\\nu_\\mu,\\nu_\\tau)$ are indeed quantum superpositions of states $(\\nu_1,\\nu_2,\\nu_3)$ with definite masses $m_i$~\\cite{Amsler:2008zz}. The simplest unitary form of the lepton mixing matrix relating flavor and mass eigenstate neutrinos is given in terms of three mixing angles $(\\theta_{12},\\theta_{13},\\theta_{23})$ and three CP-violating phases, only one of which affects (conventional) neutrino oscillations~\\cite{schechter:1980gr}. Here we consider only the effect of the mixing angles in current oscillation experiments, the sensitivity to CP violation effects remains an open challenge for future experiments~\\cite{Bandyopadhyay:2007kx,Nunokawa:2007qh}. Together with the mass splitting parameters $\\Dms \\equiv m^2_2-m^2_1$ and $\\Dma \\equiv m^2_3- m^2_1$ the angles $\\theta_{12}, \\theta_{23}$ are rather well determined by the oscillation data. In contrast, so far only upper bounds can be placed upon $\\theta_{13}$, mainly following from the null results of the short-baseline CHOOZ reactor experiment \\cite{Apollonio:2002gd} with some effect also from solar and KamLAND data, especially at low $\\Dma$ values~\\cite{Maltoni:2003da}. Here we present an update of the three-flavour oscillation analyses of Refs.~\\cite{Maltoni:2003da} and \\cite{Maltoni:2004ei}. This new analysis includes the data released this summer by the MINOS collaboration~\\cite{Adamson:2008zt}, the data from the neutral current counter phase of the SNO experiment (SNO-NCD)~\\cite{Aharmim:2008kc}, the latest KamLAND~\\cite{:2008ee} and Borexino~\\cite{Collaboration:2008mt} data, as well as the results of a recent re-analysis of the Gallex/GNO solar neutrino data presented at the Neutrino 2008 conference~\\cite{gallex-nu08:07}. In Section~\\ref{sec:lead-solar-atmosph} we discuss the status of the parameters relevant for the leading oscillation modes in solar and atmospheric neutrinos. In Section~\\ref{sec:th13} we present the updated limits on $\\theta_{13}$ and discuss the recent claims for possible hints in favour of a non-zero value made in Refs.~\\cite{Balantekin:2008zm,Escamilla:2008vq,Fogli:2008jx}. We summarize in Section~\\ref{sec:summary}. ", "conclusions": "\\label{sec:summary} In this work we have provided an update on the status of three-flavour neutrino oscillations, taking into account the latest available world neutrino oscillation data presented at the {\\em Neutrino 2008\\/} Conference. Our results are summarized in Figures~\\ref{fig:dominant-12}, \\ref{fig:dominant-23} and \\ref{fig:th13}. Table~\\ref{tab:summary} provides best fit points, $1\\sigma$ errors, and the allowed intervals at 2 and 3$\\sigma$ for the three-flavour oscillation parameters. \\begin{table}[ht]\\centering \\catcode`?=\\active \\def?{\\hphantom{0}} \\begin{tabular}{|@{\\quad}>{\\rule[-2mm]{0pt}{6mm}}l@{\\quad}|@{\\quad}c@{\\quad}|@{\\quad}c@{\\quad}|@{\\quad}c@{\\quad}|} \\hline parameter & best fit & 2$\\sigma$ & 3$\\sigma$ \\\\ \\hline $\\Delta m^2_{21}\\: [10^{-5}\\eVq]$ & $7.65^{+0.23}_{-0.20}$ & 7.25--8.11 & 7.05--8.34 \\\\[2mm] $|\\Delta m^2_{31}|\\: [10^{-3}\\eVq]$ & $2.40^{+0.12}_{-0.11}$ & 2.18--2.64 & 2.07--2.75 \\\\[2mm] $\\sin^2\\theta_{12}$ & $0.304^{+0.022}_{-0.016}$ & 0.27--0.35 & 0.25--0.37\\\\[2mm] $\\sin^2\\theta_{23}$ & $0.50^{+0.07}_{-0.06}$ & 0.39--0.63 & 0.36--0.67\\\\[2mm] $\\sin^2\\theta_{13}$ & $0.01^{+0.016}_{-0.011}$ & $\\leq$ 0.040 & $\\leq$ 0.056 \\\\ \\hline \\end{tabular} \\caption{ \\label{tab:summary} Best-fit values with 1$\\sigma$ errors, and 2$\\sigma$ and 3$\\sigma$ intervals (1 \\dof) for the three--flavour neutrino oscillation parameters from global data including solar, atmospheric, reactor (KamLAND and CHOOZ) and accelerator (K2K and MINOS) experiments.} \\end{table} \\paragraph{Acknowledgments.} This work was supported by MEC grant FPA2005-01269, by EC Contracts RTN network MRTN-CT-2004-503369 and ILIAS/N6 RII3-CT-2004-506222. We thank Michele Maltoni for collaboration on global fits to neutrino oscillation data. \\appendix" }, "0808/0808.0225_arXiv.txt": { "abstract": "{ Standard cosmology has many successes on large scales, but faces some fundamental difficulties on small, galactic scales. One such difficulty is the cusp/core problem. High resolution observations of the rotation curves for dark matter dominated low surface brightness (LSB) galaxies imply that galactic dark matter halos have a density profile with a flat central core, whereas N-body structure formation simulations predict a divergent (cuspy) density profile at the center. It has been proposed that this problem can be resolved by stellar feedback driving turbulent gas motion that erases the initial cusp. However, strong gravitational lensing prefers a cuspy density profile for galactic halos. In this paper, we use the most recent high resolution observations of the rotation curves of LSB galaxies to fit the core size as a function of halo mass, and compare the resultant lensing probability to the observational results for the well defined combined sample of the Cosmic Lens All-Sky Survey (CLASS) and Jodrell Bank/Very Large Array Astrometric Survey (JVAS). The lensing probabilities based on such density profiles are too low to match the observed lensing in CLASS/JVAS. High baryon densities in the galaxies that dominate the lensing statistics can reconcile this discrepancy, but only if they steepen the mass profile rather than making it more shallow. This places contradictory demands upon the effects of baryons on the central mass profiles of galaxies. ", "introduction": " ", "conclusions": "" }, "0808/0808.0932_arXiv.txt": { "abstract": "We present the first resolved images of the eclipsing binary $\\beta$ Lyrae, obtained with the CHARA Array interferometer and the MIRC combiner in the $H$ band. The images clearly show the mass donor and the thick disk surrounding the mass gainer at all six epochs of observation. The donor is brighter and generally appears elongated in the images, the first direct detection of photospheric tidal distortion due to Roche-lobe filling. We also confirm expectations that the disk component is more elongated than the donor and is relatively fainter at this wavelength. Image analysis and model fitting for each epoch were used for calculating the first astrometric orbital solution for $\\beta$~Lyrae, yielding precise values for the orbital inclination and position angle. The derived semi-major axis also allows us to estimate the distance of $\\beta$ Lyrae; however, systematic differences between the models and the images limit the accuracy of our distance estimate to about 15\\%. To address these issues, we will need a more physical, self-consistent model to account for all epochs as well as the multi-wavelength information from the eclipsing light curves. ", "introduction": "Interacting binaries are unique testbeds for many important astrophysical processes, such as mass and momentum transfer, accretion, tidal interaction, etc. These processes provide information on the evolution and properties of many types of objects, including low-mass black holes and neutron stars (in low-mass X-ray binaries), symbiotic binaries, cataclysmic variables, novae, etc. Although these types of objects are widely studied by indirect methods such as spectroscopy, radial velocity, and sometimes eclipse mapping, very few of them have been directly resolved because they are very close to each other and far away from us. Thus, directly imaging interacting binaries, although very challenging, will greatly help us to improve our understanding of these objects. The star $\\beta$ Lyrae (Sheliak, HD 174638, HR 7106, $V$ = 3.52, $H$=3.35) is a well known interacting and eclipsing binary that has been widely studied since its discovery in 1784 \\citep{Goodricke1785}. According to the current picture \\citep{Harmanec2002}, the system consists of a B6-8~II Roche-lobe filling mass-losing star, which is generally denoted as the donor or the primary, and an early B type mass-gaining star that is generally denoted as the gainer or the secondary. The donor, which was initially more massive than the gainer, has a current mass of about 3 \\msun, while the gainer has a mass of about 13 \\msun. It is thought that the gainer is completely embedded in a thick accretion disk with bipolar jet-like structures perpendicular to the disk, which creates a light-scattering halo above its poles. The orbit of the system is highly circular \\citep{Harmanec1993}, and is very close to edge-on \\citep{Linnell2000}. Recent RV study on the ephemeris of the system gives a period of $12.^d94$ \\citep{Ak2007}. The period is increasing at a rate of $\\sim19$ sec per year due to the high mass transfer rate, $2\\times10^{-5}\\msun$ yr$^{-1}$, of the system. The primary eclipse of the light curve (i.e., at phase 0) corresponds to the eclipse of the donor. In the $UBV$ bands, the surface of the donor is brighter than that of the gainer, and therefore the primary minimum is deeper than the secondary minimum. At longer wavelengths, however, the studies of \\citet{Jameson1976} and \\citet{Zeilik1982} suggest that the relative depth of the secondary minimum in the light curve gradually deepens and becomes deeper than the primary minimum at wavelengths longer than $3.6\\mu m$. Light curve studies and theoretical models have shown that, at the distance of 296pc \\citep{van-Leeuwen2007}, the estimated separation of the binary is only 0.92 milli-arcsecond (hereafter $mas$, $58.5\\rsun$). The angular diameter of the donor is $\\sim$0.46 mas (29.4\\rsun), and the disk surrounding the gainer is only $\\sim1$ mas across \\citep[e.g.,][]{Linnell2000, Harmanec2002}. The goal of directly imaging $\\beta$ Lyr, therefore, requires the angular resolution only achievable by today's long-baseline interferometers. Recently, \\citet{Schmitt2008} used the NPOI interferometer to image successfully the H$\\alpha$ emission of $\\beta$ Lyr, an update to the pioneering work of \\citet{Harmanec1996}. Also, radio work using MERLIN found a nebula surrounding the secondary but could not resolve its bipolar shape \\citep{Umana2000}. Despite recent progress, the individual objects of the system have not been resolved yet, putting even a simple astrometric orbit beyond our reach. In this study, we present the first resolved images of the \\betlyr system at multiple phases, obtained with the CHARA Array and the MIRC combiner. We give a brief introduction to our observations and data reduction in \\S\\ref{observations}. We present our aperture synthesis images with simple models in \\S\\ref{imaging}. In \\S\\ref{orbit} we discuss our astrometric orbit of \\betlyr and we give the outlook for future work in \\S\\ref{summary}. ", "conclusions": "" }, "0808/0808.0103_arXiv.txt": { "abstract": "% In this paper we present a number of metrics for usage of the SAO/NASA Astrophysics Data System (ADS). Since the ADS is used by the entire astronomical community, these are indicative of how the astronomical literature is used. We will show how the use of the ADS has changed both quantitatively and qualitatively. We will also show that different types of users access the system in different ways. Finally, we show how use of the ADS has evolved over the years in various regions of the world. The ADS is funded by NASA Grant NNG06GG68G. ", "introduction": "The SAO/NASA Astrophysics Data System (hereafter ADS), is a digital library and a vital source for bibliographic information in astronomy. The vast majority of astronomical researchers in the world use the ADS on a daily or near-daily basis. The use of the ADS has not only changed quantitatively but also qualitatively. Initially almost exclusively used by professional astronomers, the ADS now also has become a public service through external, general search engines (like Google, Yahoo, Microsoft Live Search and Ask.com, to name a few). In~\\citet{henneken07} we observed that up to the middle of 2004, the number of ADS users doubled on a bi-yearly basis. Since the ADS started to be indexed by general search engines, the number of incidental users has increased dramatically. However, the number of typical users (more than 10 visits per month) has continued to follow the same growth pattern. With different types of users come different types of use. A professional astronomer has different interests than an occasional user. One way of illustrating this is to look at the distribution of publication years for the literature people are interested in. We will also look at the diversity of ADS users from a geographical point of view. This will indicate whether increased Internet access actually results in an increase of ADS usage. This is particularly interesting with respect to aspects of the ``Digital Divide'' (see e.g.~\\citet{ITU07}). In the next section, we will describe the character of the data we are working with. The following section will show the results, which will then be discussed in section 4. ", "conclusions": "In terms of its audience, the ADS has not only changed quantitatively, but also qualitatively. Besides a steady growth of the ADS regular users, we observe a dramatic increase in incidental users. The ADS is per definition the gateway to online literature for scientists, used by virtually all professional astronomers on a daily basis. Since 2005 there is a growing role as a source of science education of the general public. Comparing the group of ``ADS regulars'' with the group visiting the ADS via Google Scholar shows that the obsolescence curve for the latter is fairly flat, corresponding with reading behavior by people acquainting themselves with a subject. This means Google Scholar is not the right tool for staying up-to-date with the latest events in a field. Looking at how professional astronomers use the ADS shows that the obsolescence function for them closely follows the citation rate (as a function of paper age). Although ADS usage increased in regions like the EU and the USA, the percentage of world usage has decreased for these regions. This is because the growth in World usage is mainly driven by regions with the biggest potential for growth. The density of Internet users reaches a saturation point in middle- and high-income regions at which point ADS usage increases at a slower rate. It is encouraging to see the rapid increase in Internet user density in low-income regions and a similar increase in the number of ADS users in those regions. It indicates that increased access to electronic information is being used and in this sense there is a narrowing of the ``Digital Divide'' for these regions. Whether this increased access also resulted in an increased scientific output needs further bibliometric research." }, "0808/0808.2937_arXiv.txt": { "abstract": "A framework is outlined to assess Cepheids as potential cluster members from readily available photometric observations. A relationship is derived to estimate colour excess and distance for individual Cepheids through a calibration involving recently published HST parallaxes and a cleaned sample of established cluster Cepheids. Photometric {\\it V--J} colour is found to be a viable parameter for approximating a Cepheid's reddening. The non-universal nature of the slope of the Cepheid PL relation for {\\it BV} photometry is confirmed. By comparison, the slopes of the {\\it VJ} and {\\it VI} relations seem relatively unaffected by metallicity. A new Galactic Cepheid confirmed here, GSC 03729-01127 (F6-G1 Ib), is sufficiently coincident with the coronal regions of Tombaugh 5 to warrant follow-up radial velocity measures to assess membership. CCD photometry and O--C diagrams are presented for GSC 03729-01127 and the suspected cluster Cepheids AB Cam and BD Cas. Fourier analysis of the photometry for BD Cas and recent estimates of its metallicity constrain it to be a Population I overtone pulsator rather than a Type II s-Cepheid. AB Cam and BD Cas are not physically associated with the spatially-adjacent open clusters Tombaugh 5 and King 13, respectively, the latter being much older ($\\log \\tau \\simeq 9$) than believed previously. Rates of period change are determined for the three Cepheids from archival and published data. GSC 03729-01127 and AB Cam exhibit period increases, implying fifth and third crossings of the instability strip, respectively, while BD Cas exhibits a period decrease, indicating a second crossing, with possible superposed trends unrelated to binarity. More importantly, the observed rates of period change confirm theoretical predictions. The challenges and prospects for future work in this area of research are discussed. ", "introduction": "The {\\it All Sky Automated Survey} \\citep[ASAS,][]{po00}, the {\\it Northern Sky Variability Survey} \\citep[NSVS,][]{wo04}, and {\\it The Amateur Sky Survey} \\citep[TASS,][]{dr06} have detected many new Cepheid variables through their photometric signatures, resulting in a valuable expansion of the Galactic Cepheid sample \\citep{sd04} once confirmed by spectroscopic observation. In the case of GSC 03729-01127 (TASSIV 6349369), a suspected Cepheid studied here, the variable may be an open cluster member and a potentially valuable calibrator for the Cepheid period-luminosity (PL) relation. Cepheids continue to provide the foundation for the universal distance scale, and such variables could serve as an efficient means of quantifying the extinction to Galactic and extragalactic targets. The results of the seminal Hubble Space Telescope (HST) Key Project yielded a Hubble constant of $H_0=72\\pm8$ km s$^{-1}$ Mpc$^{-1}$ \\citep{fr01}, a value supported by cosmological constraints inferred from WMAP observations \\citep{sp07}. The HST results are tied to Large Magellanic Cloud (LMC) Cepheids, which advantageously provide a large sample of common distance. However, distance estimates for the LMC exhibit an unsatisfactorly large scatter \\citep{fr01,be02}, resulting in an uncertain zero-point. Moreover there exists a difference in metallicity between LMC Cepheid variables relative to both Galactic Cepheids and those in galaxies used for calibrating secondary distance candles, the effects of which remain actively debated. \\citet{ta03}, for example, suggest that the LMC Cepheid PL relation appears to characterize short-period Cepheids as too bright relative to their Galactic counterparts, and long-period Cepheids as too faint. Conversely, \\citet{vl07} and \\citet{be07} suggest, on the basis of revised Hipparcos and newly-derived HST parallaxes, that the slopes of the PL relations ($V,I$) for Cepheids in the Galaxy and the LMC are consistent to within their cited uncertainties. Indeed, the results presented in section \\ref{extragal}, complementing in part those of \\citet{fo07}, appear to confirm ideas put forth in each of the above studies, namely that the slope of the PL relation is not universal in certain passbands, and for {\\it VJ} and {\\it VI} constructed relations, any putative difference in slope arising from metallicity effects appears negligible in comparison with other concerns and uncertainties related to extragalactic observations. Nevertheless, a consensus has yet to emerge and a resolution to the above debate may be assisted by renewed efforts towards establishing Galactic Cepheids as cluster members, a connection that provides direct constraints on Cepheid luminosities, intrinsic colours, masses, metallicities, and pulsation modes. \\citet{tu02} compiled an extensive list of suspected cluster Cepheids based upon preliminary analyses, but only a few cluster/Cepheid pairs have been studied with the necessary detail to determine the parameters of the associated clusters accurately, or to obtain the necessary radial velocity measures needed to establish membership in cases where reliable proper motions are unavailable. Efforts to discover new Galactic open clusters \\citep{al03,mo03,kro06} and Cepheid variables \\citep{po00,wo04,dr06} have resulted in a welcome increase to the number of suspected cluster Cepheids. This paper outlines a framework to assess the viability of such cases efficiently, with an intent to highlight cases requiring further attention and focus. Section \\ref{framework} develops a relationship to estimate colour excesses and distances for individual Cepheids from several photometric parameters. Section \\ref{cepheids} presents CCD photometry, spectroscopic results, and O--C analyses for the suspected cluster Cepheids BD Cas \\citep{ts66,tu02}, AB Cam \\citep{vb57,ts66}, and a new Galactic Cepheid confirmed here, GSC 03729-01127. Distances, colour excesses, and ages are also derived for the associated open clusters Tombaugh 5 and King 13 from 2MASS photometry \\citep{cu03}. ", "conclusions": "The relationships highlighted in section 3 yield reliable parameters when investigating short period Cepheids ($P\\le11^{\\rm d}$). Parameters determined for longer period Cepheids from such relations are less certain, primarily because of an absence of mid-to-long period calibrators needed to identify a unique set of co-efficients consistent over a broad period baseline. At present $\\ell$ Car is the only established long-period calibrator (parallax) used in deriving the co-efficients. A further drawback of the analysis rests in the adopted parameters for the calibrating clusters, which exhibit an unsatisfactory amount of scatter in the literature and more recent analyses \\citep[e.g.,][]{an07,ho03}. Refining distance estimates to the calibrating set of clusters by means of deep CCD photometry, analagous to the impressive results from the CFHT Open Cluster Survey \\citep{ka01a,ka01b}, is a priority in moving forward. The framework is also tied to the HST sample of Cepheids with parallaxes and field reddenings established by \\citet{be07}. It is noted that the parallax measures for RT Aur and Y Sge differ significantly between HST and Hipparcos \\citep[][see their table 1]{vl07}. The framework outlined here should permit an efficient investigation of suspected cluster Cepheids \\citep{tu02}, including objects uncovered by cross correlations between newly discovered Cepheids in the ASAS and TASS with open cluster databases (i.e., WebDA). A potential goal is an expansion of the sample of cluster Cepheids, with particular emphasis on long period cluster Cepheids. Of equal importance, however, is the task of purging line-of-sight coincidences from current lists promulgating the literature, something that is particularly acute given that high-precision data are needed to address the question of the universality of the PL relation. Four longer term objectives exist. First is to use the new relations to determine Galactic parameters and map interstellar extinction. Second, is to establish mean photometry for an entire calibrating set. Third, with regard to the universality of the PL relation and establishing long-period Cepheid calibrators, realistically it will be the highly anticipated results from the GAIA mission \\citep{ct06}, a next generation follow-up to the Hipparcos mission, that will provide the large and unbiased sample of Cepheid parallaxes needed to advance our knowledge of the field. In conjunction with a cleaned sample of cluster Cepheids, it should lead to a proper refinement of the relations outlined in section \\ref{framework}, and, consequently, the realization of the outlined objectives. Fourth is the longer term prospect of conducting extragalactic surveys using JWST \\citep{ga06} to determine the distances and extinction to, and within [equation (\\ref{eqn4})], higher redshift galaxies. The aperture size and infrared sensitivity of the telescope will permit deeper sampling of extragalactic Cepheids, especially since the variables are substantially brighter in the infrared than the optical and the diminishment in flux from reddening is comparitively less. \\subsection*{ACKNOWLEDGEMENTS} We are indebted to the following individuals and groups who helped facilitate the research: Alison Doane and the staff of the Harvard College Observatory Photographic Plate Stacks, Charles Bonatto for useful discussions on taking advantage of data from the 2MASS survey, Pascal Fouqu\\'{e}, Laszlo Szabados and Leonid Berdnikov, whose comprehensive work on evolutionary trends in Cepheid variables was invaluable in our analysis, Arne Henden and the staff at the AAVSO, Dmitry Monin, Les Saddelmeyer, and the rest of the staff of the Dominion Astrophysical Observatory, Doug Welch who maintains the McMaster Cepheid Photometry and Radial Velocity Archive, the staff at la Centre de Donn\\'{e}es astronomiques de Strasbourg, and Carolyn Stern Grant and the staff at the Astrophysics Data System (ADS). Reviews on Cepheids by Michael Feast, Donald Fernie, and Nick Allen were useful in the preparation of this work. Lastly, we extend a special thanks to Sandra Hewitt for her exceptional kindness in accommodating visiting astronomers to the Harvard Plate Stacks." }, "0808/0808.0759_arXiv.txt": { "abstract": "The comparison of the black hole mass function (BHMF) of active galactic nuclei (AGN) relics with the measured mass function of the massive black holes in galaxies provides strong evidence for the growth of massive black holes being dominated by mass accretion. We derive the Eddington ratio distributions as functions of black hole mass and redshift from a large AGN sample with measured Eddington ratios given by Kollmeier et al. We find that, even at the low mass end, most black holes are accreting at Eddington ratio $\\lambda\\sim0.2$, which implies that the objects accreting at extremely high rates should be rare or such phases are very short. Using the derived Eddington ratios, we explore the cosmological evolution of massive black holes with an AGN bolometric luminosity function (LF). It is found that the resulted BHMF of AGN relics is unable to match the measured local BHMF of galaxies for any value of (constant) radiative efficiency $\\eta_{\\rm rad}$. Motivated by Volonteri, Sikora \\& Lasota's study on the spin evolution of massive black holes, we assume the radiative efficiency to be dependent on black hole mass, i.e., $\\eta_{\\rm rad}$ is low for $M_{\\rm bh}<10^8{\\rm M}_\\odot$ and it increases with black hole mass for $M_{\\rm bh}\\ge 10^8{\\rm M}_\\odot$. We find that the BHMF of AGN relics can roughly reproduce the local BHMF of galaxies if $\\eta_{\\rm rad}\\simeq0.08$ for $M_{\\rm bh}<10^8{\\rm M}_\\odot$ and it increases to $\\ga 0.18$ for $M_{\\rm bh}\\ga10^9{\\rm M}_\\odot$, which implies that most massive black holes ($\\ga 10^9{\\rm M}_\\odot$) are spinning very rapidly. ", "introduction": "It is believed that quasars are powered by accretion on to massive black holes, and the growth of the massive black holes could be governed by mass accretion in quasars. The massive black holes (AGN relics) should be present in the centres of galaxies. Thus, the luminosity functions (LF) of active galactic nuclei (AGN) provide important clues on the growth of massive black holes. It is indeed found that most nearby galaxies contain massive black holes at their centres, and a tight correlation is revealed between central massive black hole mass and the velocity dispersion of the galaxy \\citep{fm00,g00}. The black hole mass is also found to be tightly correlated with the luminosity of the spheroid component of its host galaxy \\citep*[e.g.,][]{m98,mh03}. These correlations of the black hole mass with velocity dispersion/host galaxy luminosity were used to derive the mass functions of the central massive black holes in galaxies \\citep*[e.g.,][]{yt02,m04,t06,g07}. On the other hand, the black hole mass function (BHMF) of AGN relics can also be calculated by integrating the continuity equation of massive black hole number density on the assumption of the growth of massive black holes being dominated by mass accretion, in which the activity of massive black holes is described by a LF of AGN \\citep*[e.g.,][]{c71,soltan82,ct92,sb92,m04,s04,t06}. Such calculations on the cosmological evolution of massive black holes were usually carried out by adopting two free parameters: the radiative efficiency $\\eta_{\\rm rad}$ and the Eddington ratio $\\lambda$ for AGN. The derived BHMF of AGN relics in this way is required to match that estimated either from velocity dispersion or the luminosity of the spheroid of its host galaxy, which always leads to $\\lambda\\sim 1$, i.e., almost all AGN are required to be accreting close to the Eddington limit \\citep*[e.g.,][]{yt02,m04,s04,t06}. In the last decade, several approaches for measuring the masses of the central black holes in AGN have been developed, in which the reverberation mapping may be the most effective one \\citep{p93,k00}. Using the tight correlation between the size of the broad-line region and the optical luminosity established with the reverberation mapping method for a sample of AGN, the black hole masses of AGN can be easily estimated from their optical luminosity and width of broad emission line. The Eddington ratios for thousands of AGN were estimated with the analyses of the Sloan Digital Sky Survey (SDSS) by \\citet{md04}, which indicate that the mean Eddington ratio $L_{\\rm bol}/L_{\\rm Edd}\\simeq 0.1$ at $z\\sim 0.2$ to $\\simeq0.4$ at $z\\sim 2$. \\citet{w04} also derived the Eddington ratio distribution for a sample of $\\sim$~500 AGN with redshifts $0\\la z\\la 5$. As pointed by \\citet{k06}, both these derived Eddington ratios are heavily weighted towards high-luminosity objects due to the limited sensitivity of SDSS. \\citet{k06} estimated the Eddington ratios of AGN discovered in the AGN and Galaxy Evolution Survey (AGES), which is more sensitive than the SDSS \\citep{k04}. The derived Eddington ratio distribution at {\\it fixed luminosity} is well described by a single lognormal distribution peaked at $\\sim 0.25$ independent of redshift and luminosity \\citep*[see][for the details]{k06}. In this work, the Eddington ratio distribution at {\\it fixed luminosity} given by \\citet{k06} is converted to that at {\\rm fixed black hole mass} by using an AGN LF. We integrate the continuity equation for black hole number density adopting the derived Eddington ratio distributions of AGN to calculate the BHMF of AGN relics at different redshifts $z$, which is different from a free parameter $\\lambda$ adopted in most previous works. The resultant BHMF of AGN relics is constrained by those estimated from the galaxy LFs \\citep{m04,t06}. The conventional 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 work. ", "conclusions": "As in the most previous works, we implicitly assume that the black hole growth is dominated by mass accretion in bright AGN, while some inactive black holes may still be accreting gases, though their mass accretion rates are very low. If the duration of the accretion in these objects is as long as the Hubble timescale, they can accrete sufficient mass comparable with that accumulated in bright AGN phases, as the AGN phase is much shorter than the Hubble timescale. It is believed that the advection dominated accretion flows (ADAFs) are present in those objects, which are very hot and radiate mostly in hard X-ray bands \\citep{ny94}. They are very difficult to be detected due to low luminosity, unless those in the nearby Universe. \\citet{cao05} suggested that the accretion of such low-luminosity objects can be constrained by the hard X-ray background, though the emission from most of these individuals cannot be detected by any facilities now. It was found that less than $\\sim 5$ per cent of the local black hole mass density was accreted during the ADAF phases, which will be even lower if the Compton-thick AGN are included \\citep*[see][for the details]{cao07}. \\citet{h06} considered the distribution of local supermassive black hole Eddington ratios and accretion rates, accounting for the dependence of radiative efficiency and bolometric corrections on the accretion rate. They also found that black hole mass growth was dominated by AGN phase, and not by the radiatively inefficient low accretion rate phase in which most local supermassive black holes are currently observed. The main difference of this work from the previous works is that the Eddington ratio distributions are derived from an AGN sample with measured Eddington ratios \\citep{k06}. The Eddington ratio distributions for fixed black hole mass derived in our work approximate to the lognormal distribution (see Fig. \\ref{fig1}), and the mean Eddington ratios are in the range of $\\sim 0.1-0.3$ varying with black hole mass and redshift (see Fig. \\ref{fig2}). For most cases, the mean Eddington ratios peak at $\\sim 10^8{\\rm M}_\\odot$, and then decline with increasing black hole mass. Even at the low mass end, most black holes are accreting at $\\lambda\\sim0.2$, which implies that the objects accreting at extremely high rates should be rare or such phases are very short. It was suggested that the radiative efficiency $\\eta_{\\rm rad}$ declines for a slim accretion disc provided the mass accretion rate is sufficiently high due to the photon trapping effort \\citep[e.g.,][]{a88,b78,w99}. \\citet{w00}'s calculations on the slim discs showed that the radiative efficiency will not deviate significantly from that for standard thin discs if $L_{\\rm bol}/L_{\\rm Edd}\\la 2$, which implies that the present adopted radiative efficiency independent of Eddington ratio $\\lambda$ is indeed a good assumption. There is only one free parameter $\\eta_{\\rm rad}$ in our calculations for the cosmological evolution of massive black holes. We find that the resulted BHMF of AGN relics is unable to reproduce the measured local BHMF for any value of $\\eta_{\\rm rad}$ adopted, provided the radiative efficiency $\\eta_{\\rm rad}$ is independent of black hole mass, as treated in previous works \\citep*[e.g.,][]{yt02,m04,t06}. The mean Eddington ratios adopted in our calculations are derived from an AGN sample, which are in the range of $\\sim 0.1-0.3$. Thus, it is not surprising that the local BHMFs cannot be reproduced by our calculations with any constant radiative efficiency, because the Eddington ratio $\\lambda\\sim 1$ is usually required in order to let the resulted BHMF match the local one in those works. In this work, we use two different corrections (either luminosity-independent or luminosity-dependent) for the Compton-thick AGN (see Sect. 3 for the details), and find that the final results are quite similar (see Figs. \\ref{fig3} and \\ref{fig4}). We also use the hard X-ray LF derived from an AGN sample at high redshifts by \\citet{s08} in stead of the bolometric LF of \\citet{h07} in the calculations. It is found that the main results of this work change very little and the main conclusion is not altered. \\citet{v07} studied on how the accretion from a warped disc influences the evolution of black hole spins and concluded that within the cosmological framework, one indeed expects most supermassive black holes in elliptical galaxies to have on average higher spin than black holes in spiral galaxies, where random, small accretion episodes (e.g., tidally disrupted stars, accretion of molecular clouds) might have played a more important role. Thus, we tentatively adopt a $M_{\\rm bh}$-dependent radiative efficiency (see Eq. \\ref{etarad}), in which $\\eta_{\\rm rad}$ remains constant for $M_{\\rm bh}\\le 10^8$ and increases with black hole mass for $M_{\\rm bh}> 10^8$. This $M_{\\rm bh}$-dependent radiative efficiency is qualitatively consistent with the results of \\citet{v07}. It is found that the measured BHMFs can be fairly well reproduced by our model calculations with this $M_{\\rm bh}$-dependent radiative efficiency (see Figs. \\ref{fig3} and \\ref{fig4}), which require $\\eta_{\\rm rad}\\ga 0.18$ for $M_{\\rm bh}\\ga 10^9{\\rm M}_\\odot$. This provides evidence for most massive black holes being spinning very rapidly. It is interesting to find that $a\\simeq 0.9$ is also required by the fitting of the residual hard X-ray background with the emission from the ADAFs in the low-luminosity objects \\citep{cao07}. Our calculations can be improved if the mean spin parameter $a$ as a function of black hole mass is available from the work within the cosmological framework, which is beyond the scope of present work. { In our present calculations of the black hole evolution, the black hole mergers have been neglected. \\citet{s07} assessed the importance of the black hole mergers on the evolution of the BHMF. They found that the impact of black hole mergers on the cosmological evolution of BHMF may probably be small compared with black hole accretion processes, while its impact on the black hole spin evolution may be important \\citep*[e.g.,][]{wc95,hb03,v05,v07}. The effect of black hole mergers increases the number density of very massive black holes \\citep{s07}, which implies that the radiative efficiencies for very massive active black holes should be higher than the present values if black hole mergers are included in our calculations. This strengthens our conclusion that most massive black holes are spinning very rapidly. }" }, "0808/0808.3793_arXiv.txt": { "abstract": "The apparent spectral evolution observed in the steep decay phase of many GRB early afterglows raises a great concern of the high-latitude ``curvature effect'' interpretation of this phase. However, previous curvature effect models only invoked a simple power law spectrum upon the cessation of the prompt internal emission. We investigate a model that invokes the ``curvature effect'' of a more general non-power-law spectrum and test this model with the Swift/XRT data of some GRBs. We show that one can reproduce both the observed lightcurve and the apparent spectral evolution of several GRBs using a model invoking a power-law spectrum with an exponential cut off. GRB 050814 is presented as an example. ", "introduction": "} Most of the early X-Ray afterglows detected by Swift (Gehrels et al. 2004) show a steep decay phase around 100$\\sim$1000 seconds after the burst trigger (Tagliaferri et al. 2005). The main characteristics of this steep decay phase include the following. (1) It connects smoothly to the prompt $\\gamma$-ray light curve extrapolated to the X-ray band, suggesting that it is the ``tail'' of the prompt emission (Barthelmy et al. 2005, O'Brien et 2006, Liang et al 2006). (2) The decay slope is typically $3 \\sim 5$ when choosing the GRB trigger time as the zero time point $t_0$ (Tagliaferri et al. 2005; Nousek et al. 2006; Zhang et al 2006). (3) The time-averaged spectral index of the steep decay phase is much different from that of the later shallow decay phase, indicating that it is a distinct new component that is unrelated to the conventional afterglow components (Zhang et al 2006; Liang et al. 2007). (4) Strong spectral evolution exists in about one third of the bursts that have a steep decay phase (Zhang et al. 2007, hereafter ZLZ07; Butler \\& Kocevski 2007; Starling et al. 2008). All these features suggest that the steep decay phase holds the key to understand the connection between the prompt emission (internal) phase and the traditional afterglow (external) phase. Any proposed model (see M\\'esz\\'aros 2006; Zhang 2007 for reviews) should be able to explain these features. The so called ``curvature effect'', which accounts for the delayed photon emission from high latitudes with respect to the line of sight upon the abrupt cessation of emission in the prompt emission region (Fenimore et al. 1996; Kumar \\& Panaitescu 2000; Dermer 2004; Dyks et al. 2005; Qin 2008a), has been suggested to play an important role in shaping the sharp flux decline in GRB tails (Zhang et al. 2006; Liang et al. 2006; Wu et al. 2006; Yamazaki et al. 2006). In the simplest model, it is assumed that the instantaneous spectrum at the end of the prompt emission is a simple power law with a spectral index $\\beta$. The predicted temporal decay index of the emission is (with the convention $F_{\\nu} \\propto t^{-\\alpha} \\nu^{-\\beta}$) \\begin{equation} \\alpha=2+\\beta~, \\label{curvature-1} \\end{equation} if the time origin to define the $\\log-\\log$ light curve, $t_0$, is taken as the beginning of the last emission episode before the cessation of emission. Adopting a time-averaged $\\beta$ in the tails, Liang et al. (2006) found that Eq.(\\ref{curvature-1}) is generally valid. The strong spectral evolution identified in a group of GRB tails (ZLZ07) apparently violates Eq.(\\ref{curvature-1}), which is valid only for a constant $\\beta$. ZLZ07 then investigated a curvature effect model by assuming a structured jet with varying $\\beta$ at different latitudes and that the line of sight is near the jet axis\\footnote{Notice that this structured jet model is different from the traditional one that invokes an angle-dependent energy/Lorentz factor, but not the spectral index (Zhang \\& M\\'esz\\'aros 2002; Rossi et al. 2002).}. One would then expect that Eq.(\\ref{curvature-1}) is roughly satisfied, with both $\\alpha$ and $\\beta$ being time-dependent. ZLZ07 found that this model does not fit the data well. These facts do not rule out the curvature effect interpretation of GRB tails, however. This is because the instantaneous spectrum upon the cessation of prompt emission may not be a simple power law. If the spectrum has a curvature, as the emission from progressively higher latitudes reach the observer, the XRT band is sampling different segments of the intrinsic curved spectrum (Fig.1). This would introduce an apparent spectral evolution in the decaying tail. The main goal of this paper is to test this more general curvature effect model using the available Swift XRT data. ", "conclusions": "We have successfully modeled the lightcurve and spectral evolution of the X-ray tail of GRB050814 using the curvature effect model of a cutoff power law spectrum with an exponential cutoff ($k=1$). It has been discussed in the literature (e.g. Fan \\& Wei 2005; Barniol-Duran \\& Kumar 2008) that the GRB central engine may not die abruptly, and that the observed X-ray tails may reflect the dying history of the central engine. If this is indeed the case, the strong spectral evolution in the X-ray tails would demand a time-dependent particle acceleration mechanism that gives a progressively soft particle spectrum. Such a behavior has not been predicted by particle acceleration theories. Our results suggest that at least for some tails, the spectral evolution is simply a consequence of the curvature effect: the observer views emission from the progressively higher latitudes from the line of sight, so that the XRT band is sampling the different segments of a curved spectrum. This is a simpler interpretation. The phenomenology of the X-ray tails are different from case to case (ZLZ07). We have applied our model to some other clean X-Ray tails, such as GRB050724, GRB080523, and find that they can be also interpreted by this model. Some other tails have superposed X-ray flares, making a robust test of the model difficult. A systematic survey of all the data sample is needed to address what fraction of the bursts can be interpreted in this way or they demand other physically distinct models (e.g. Barniol-Duran \\& Kumar 2008; Dado et al. 2008). This is beyond the scope of this Letter." }, "0808/0808.3046_arXiv.txt": { "abstract": "s{We present a powerful method for exploring various processes in the presence of strong external fields and matter. The method implies utilization of the exact solutions of the modified Dirac equations which contain the effective potentials accounting for the influences of external electromagnetic fields and matter on particles. We briefly discuss the basics of the method and its applications to studies of different processes, including a recently proposed new mechanism of radiation by neutrinos and electrons moving in matter (the spin light of the neutrino and electron). In view of a recent ``prediction'' of an order-of-magnitude change of the muon lifetime under the influence of an electromagnetic field of a CO$_2$ laser, we revisit the issue and show that such claims are nonrealistic.} ", "introduction": "The problem of particles' interactions under the influence of external electromagnetic fields and matter is one of the important topics in particle physics. Besides the possibility for better visualization of fundamental properties of particles and their interactions when they are influenced by external conditions, the interest to this problem is also stimulated by astrophysical and cosmological applications, where strong electromagnetic fields and dense matter may play important roles. There are well established methods for such kind of investigations that have a long-standing history. In particular, the method of exact solutions of quantum equations, which is based on a Furry representation \\cite{FurPR51} of QED, is widely used in studies of particles' interactions in external electromagnetic fields. In this technique, the evolution operator $U_{F}(t_1, t_2)$, which determines the matrix element of the process, is presented in the usual form \\begin{equation} U_{F} (t_1, t_2)= T exp \\left[-i \\int \\limits_{t_1}^{t_2}j_{\\mu}(x) A^{\\mu}{d}x \\right], \\end{equation} where $A_{\\mu}(x)$ is the quantized part of the potential corresponding to the radiation field, which is accounted for within the perturbation-series techniques. At the same time, the electron (a charged particle) current is presented as \\begin{equation} j_{\\mu}(x)={e \\over 2}\\left[\\overline \\Psi_e \\gamma _{\\mu}, \\Psi_e \\right], \\end{equation} where $\\Psi_e$ are the exact solutions of the Dirac equation for an electron in the presence of an external electromagnetic field given by the classical non-quantized potential $A_{\\mu}^{ext}(x)$: \\begin{equation}\\label{D_eq_QED} \\left\\{ \\gamma^{\\mu}\\left(i\\partial_{\\mu} -eA_{\\mu}^{cl}(x)\\right) - m_e \\right\\}\\Psi_e (x)=0. \\end{equation} Note that within this approach the interaction of charged particles with an external electromagnetic field is taken into account exactly while the radiation field is allowed for by perturbation-series expansion techniques. That is why the method discussed is often referred to as the ``method of exact solutions''. The detailed discussion on foundations of this method and its applications to different processes such as, for instance, the synchrotron radiation by an electron in magnetic fields can be found in~\\cite{SokTerSynRad68,ritus87,nikishov87}. Many results, which ate very important for astrophysical applications, have been obtained within the discussed method when considering the neutron beta-decay in a constant magnetic field. These studies have been started in~\\cite{Kor64TerLysKor65}. Short reviews on the studies of beta-decay and the related cross symmetric processes in strong magnetic fields can be found in~\\cite{ShiStuP05KouStuPRC05}. ", "conclusions": "" }, "0808/0808.1872_arXiv.txt": { "abstract": "Planetary nebulae (PN) are an essential tool in the study of the chemical evolution of the Milky Way and galaxies of the Local Group, particularly the Magellanic Clouds. In this work, we present some recent results on the determination of chemical abundances from PN in the Large and Small Magellanic Clouds, and compare these results with data from our own Galaxy and other galaxies in the Local Group. As a result of our continuing long term program, we have a large database comprising about 300 objects for which reliable abundances of several elements from He to Ar have been obtained. Such data can be used to derive constraints to the nucleosynthesis processes in the progenitor stars in galaxies of different metallicities. We also investigate the time evolution of the oxygen abundances in the SMC by deriving the properties of the PN progenitor stars, which include their masses and ages. We have then obtained an age-metallicity relation taking into account both oxygen and [Fe/H] abundances. We show that these results have an important consequence on the star formation rate of the SMC, in particular by suggesting a star formation burst in the last 2-3 Gyr. ", "introduction": "Planetary nebulae (PN) are an essential tool in the study of the chemical evolution of the Milky Way and galaxies of the Local Group, particularly the Magellanic Clouds (see for example \\cite[Maciel et al. 2006a]{maciel2006a}, \\cite[Richer \\& McCall 2006] {richer}, \\cite[Buzzoni et al. 2006] {buzzoni}, \\cite[Ciardullo 2006)] {ciardullo}. As the offspring of stars within a reasonably large mass bracket (0.8 to about 8 solar masses), PN encompass an equally large age spread, as well as different spatial and kinematic distributions. They usually present bright emission lines and can be easily distinguished from other emission line objects, so that their chemical composition and spatiokinematical properties are relatively well determined. In this work, we present some recent results on the determination of chemical abundances from PN in the Large and Small Magellanic Clouds, and compare these results with data from our own Galaxy and other galaxies in the Local Group. We also investigate the time evolution of the oxygen abundances in the SMC by deriving an age-metallicity relation for this object. ", "conclusions": "" }, "0808/0808.3336_arXiv.txt": { "abstract": "We present the atlas of protoplanetary disks in the Orion Nebula based on the ACS/WFC images obtained for the \\emph{HST Treasury Program on the Orion Nebula Cluster}. The observations have been carried out in 5 photometric filters nearly equivalent to the standard $B, V, H\\alpha, I,$ and $ z$ passbands. Our master catalog lists 178 externally ionized proto-planetary disks (\\emph{proplyds}), 28 disks seen only in absorption against the bright nebular background (\\emph{silhouette disks}, 8 disks seen only as dark lanes at the midplane of extended polar emission (\\emph{bipolar nebulae} or \\emph{reflection nebulae}) and 5 sources showing jet emission with no evidence of neither external ionized gas emission nor dark silhouette disks. Many of these disks are associated with jets seen in $H\\alpha$ and circumstellar material detected through reflection emission in our broad-band filters; approximately 2/3 have identified counterparts in x-rays. A total of 47 objects (29 proplyds, 7 silhouette disks, 6 bipolar nebulae, 5 jets with no evidence of proplyd emission or silhouette disk) are new detections with HST. We include in our list 4 objects previously reported as circumstellar disks which have not been detected in our HST/ACS images either because they are hidden by the bleeding trails of a nearby saturated bright star or because of their location out of the HST/ACS Treasury Program field. Other 31 sources previously reported as extended objects do not harbor a stellar source in our HST/ACS images. We also report on the detection of 16 red, elongated sources. Their location at the edges of the field, far from the Trapezium Cluster core ($\\gtrsim 10'$), suggests that these are probably background galaxies observed through low extinction regions of the Orion Molecular Cloud OMC-1. ", "introduction": "The Orion Nebula (M42, NGC 1976) is a unique laboratory for studying the physical processes related to star and planet formation. It harbors one of the richest and youngest clusters (Orion Nebula Cluster, ONC)\\ in the solar neighborhood, spanning the full spectrum of stellar and sub-stellar masses down to a few Jupiter masses \\citep{Luca00}. In 1979 several compact photoionized knots were firstly detected in the central region of the Orion Nebula as emission-line sources \\citep{Laqu79}, and then important follow-up studies were made in radio \\citep{Gara87, Chur87} and via emission-line spectroscopy \\citep{Meab88, Meab93, Mass93}. Since the early 1990's, \\emph{Hubble Space Telescope} (HST) observations of the ONC have been fundamental for clarifying the main characteristics of these young stellar objects (YSO) and their accretion disks. After the pioneering surveys of \\citet{Odel93} and \\citet{Pros94}, performed with the spherically-aberrated WF/PC, \\citet{Odel94} used WFPC2 to discover several externally ionized proto-planetary disks (\\emph{proplyds}), as well as a number of disks seen only in absorption against the bright nebular background (\\emph{silhouette disks}), both rendered visible by their location in or near the core of the \\ion{H}{2} region. Following this discovery, other HST programs have increased the number of known objects \\citep{Odel96, McCa96, Ball98, Ball00}. \\citet{Odel01} and \\citet{Smit05}, targeting areas out of the core, showed that these systems are ubiquitous across the Great Orion Nebula. So far a total of $\\sim$ 200 silhouette disks and bright proplyds has been revealed by the HST observations of the Orion Nebula, the large majority through narrow-band filters centered on the H$\\alpha$ $\\lambda 6563$ emission lines, and occasionally through filters centered on the [N II] $\\lambda 6583$, [O I] $\\lambda 6300$, [O III] $\\lambda 5007$ and [S II] $\\lambda 6717+6731$ lines. In this paper we present an atlas of multi-color observations of circumstellar disks and resolved circumstellar emission obtained with the Wide Field Channel of the Advanced Camera for Surveys (ACS/WFC). These images are part of the \\emph{HST Treasury Program on the Orion Nebula Cluster} (Cycle 13, GO Program 10246, P.I. M. Robberto), aimed at measuring with high precision the main stellar parameters of the cluster members. For this reason, the Treasury Program used broad-band filters to obtain the most accurate photometry of each source, together with H$\\alpha$ narrow-band images to address the presence of circumstellar emission that may contaminate the photometry and the point spread function of the broad-band data. The combination of broad-band and narrow-band images opens a new window on the study of disks in the OMC. It also makes it possible to detect disks where the nebular background is too faint, thanks to the light of the central stars reflected by the circumstellar material at the disk's polar regions (reflection nebulae). After a brief description of the observations (\\S 2), we present the new ACS/WFC images of all circumstellar disks (\\S 3). We then provide a complete catalogue of circumstellar disks in the Orion Nebula, including also the few disks that were not detected in our programs for a variety of reasons (\\S 4). Finally, after a brief description of the new proplyds, silhouette disks, and bipolar nebulae (\\S 5, \\S 6) we present the images of 16 red, elongated, and diffuse objects which most probably represent galaxies seen through the background curtain provided by the Orion Molecular Cloud (OMC-1, \\S 7). A few remarkable objects are being investigated and will be discussed in separate papers (see e.g. \\citet{Robb08b} on 124-132). ", "conclusions": "In this paper we have shown the HST/ACS images of the 178 proplyds, 28 disks seen only in silhouette, 8 reflection nebulae without external ionized plasma, 5 jets without neither external ionized plasma nor silhouette disk and 16 other extended objects observed by the HST Treasury Program on the Orion Nebula Cluster. For every object we have reported all the images taken through the 5 photometric filters used by the HST Treasury Program (F435W, F555W, F658N, F775W, F850LP). The fact that most of these objects are associated to X-rays sources observed by the Chandra Orion Ultradeep Project is particularly interesting, since high energy photons could play an important role in the star and planet formation processes. Among all the objects reported, 63 have been discovered by these images: 29 proplyds, 7 silhouette disks, 6 reflection nebulae with no external ionized plasma, 5 jets with no external ionized plasma or silhouette disk, and 16 other elongated object. Searching in the literature we found that 4 objects previously reported as circumstellar disks have not been detected by HST/ACS images either because hidden by the saturation bleeding trails of a close bright star or because located out of HST/ACS Treasury Program field of view. For other 30 sources previously reported as extended objects HST/ACS images reveal no circumstellar emission around them. A brief description of all the newly discovered proplyds, disks seen only in silhouette and reflection nebulae with no external ionized plasma has been carried out in \\S 5 and \\S 6. Finally, we have discussed possible interpretations for the nature of the 16 extended objects. Because of their location far from the Trapezium Cluster ($\\gtrsim 10'$) and because of their red color, they are probably background galaxies reddened by the Orion Molecular Cloud OMC-1, but the alternative hypothesis of reflection nebulae turned on by red pre-main-sequence stars cannot be ruled out by our observations only. \\\\ \\\\ We wish to thank the referee, William Henney, for his useful comments that greatly improved the manuscript. Support for program 10246 was provided by NASA through a grant from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. The work of LR at STScI was done under the auspices of the STScI Summer Student Program." }, "0808/0808.0515_arXiv.txt": { "abstract": "{Using observed GALEX far-ultraviolet (FUV) fluxes and VLA images of the 21-cm \\HI\\ column densities, along with estimates of the local dust abundances, we measure the volume densities of a sample of actively star-forming giant molecular clouds (GMCs) in the nearby spiral galaxy M83 on a typical resolution scale of 170 pc.\\\\ Our approach is based on an equilibrium model for the cycle of molecular hydrogen formation on dust grains and photodissociation under the influence of the FUV radiation on the cloud surfaces of GMCs.\\\\ We find a range of total volume densities on the surface of GMCs in M83, namely 0.1 - 400 cm$^{-3}$ inside $R_{25}$, 0.5 - 50 cm$^{-3}$ outside $R_{25}$. Our data include a number of GMCs in the \\HI\\ ring surrounding this galaxy. Finally, we discuss the effects of observational selection, which may bias our results.} ", "introduction": "This paper aims to measure the total gas densities in a sample of giant molecular clouds (GMCs) in the nearby spiral M83 (NGC 5236), using a method initially proposed by \\citet{1997ApJ...487..171A}. This method is based on the simple (but unavoidable) fact that ultraviolet photons from young, newly-formed stars will react back on the surrounding parent GMCs, dissociating the molecular gas on their surfaces and turning the (virtually invisible) \\Htwo\\ into its easily-detected atomic form. The motivation for this approach was provided by the discovery of \\citet{1986Natur.319..296A} that the \\HI\\ delineating the spiral arms in a nearby galaxy showed a large-scale morphology that was more consistent with photodissociation near the \\HII\\ regions than it was with compression of the \\HI\\ farther upstream in the spiral shock. The presence of atomic hydrogen is then indicative of a photodissociation region (PDR). The method was first applied in some detail to M101 by \\citet{2000ApJ...538..608S} and more recently to M81 by \\citet{2008ApJ...673..798H}. The large, nearby spiral M83 has been a frequent target of searches for molecular gas using the CO(1-0) spectral line, which is easily detectable in M83. However, such studies have only recently begun to be carried out with sufficient linear (spatial) resolution to discern differences in the location of molecular gas and the emissions from young, hot stars. For example, \\citet{1991ApJ...381..130L} describe the presence of molecular (CO) emission ~300 pc downstream from M83's eastern spiral arm dust lane, detected with a $5\\arcsec\\ \\times 12\\arcsec$ beam. \\citet{1999ApJ...513..720R} find that the CO emission is spatially separated from the dust lane as well as the young stars on scales of a few hundred parsec. They suggest UV heating, cosmic-ray heating or a two-component molecular phase to explain this morphology. Their observations show features that show similarity to the largest GMCs in the Milky Way with masses on the order of several millions of solar masses. \\citet{1993ApJ...414...98W} explore the CO content in the central kiloparsec of M83 and infer a low-density component ($n_{H_2} \\lesssim 10^3-10^4 \\rm{cm}^{-3}$) and a warm (above 50 K), high-density component ($n_{H_2} \\gtrsim 10^4-10^5 \\rm{cm}^{-3}$). \\citet{2002AJ....123.1892C} study CO(1-0), CO(2-1) and neutral gas in M83, observing a strong truncation of the molecular disk at 6\\arcmin\\ accompanied by a warped atomic outer disk. Using the conventional assumptions about how to convert CO surface brightness to \\Htwo\\ column density, they conclude that roughly 80\\% of the total gas mass in M83 is \\Htwo. \\correction{The} \\HI\\ 21 cm and CO surface brightness are found to be correlated, but with a complex pattern of offsets. They speculate that the temperature of the CO gas is $>$ 20K in the nucleus and $<$ 7K in the outer disk. \\citet{2004A&A...413..505L} give a full overview of previous observations and present full CO (1-0) and CO (2-1) maps based on thousands of telescope pointings, but with a modest spatial resolution of $\\sim 1$ kpc. They conclude that the molecular gas spiral arms mostly trace the dust lanes. They expect the \\Htwo\\ mass to dominate that of \\HI\\ within 7.3 kpc of the center. At their linear resolution, CO and \\HI\\ emissions are correlated strongly within the optical disk. \\citet{2005A&A...441..491V} look into various tracers of star formation, including Polycyclic Aromatic Hydrocarbon (PAH) emission lines, another potential indicator of PDRs. Various combinations of these lines prove to be good tracers of star forming regions in M83. They are found predominantly in the spiral arms. We studied the occurrence of PAHs near PDRs in M81 \\citep{2008ApJ...673..798H} previously and confirmed that they are found near the star forming regions in almost all cases. Similar M83 data was not available at this time. \\begin{figure*}[t!] \\centering \\includegraphics[width=17cm]{locplot} \\caption{The location of the candidate PDRs are shown with the GALEX FUV image in the background. Not all FUV sources are visible on this image, but all sources are sources of FUV radiation. The inner circle signifies a 1 kpc galactocentric radius, the outer one signifies $R_{25}$.} \\label{fig:locplot} \\end{figure*} \\citet{2005ApJ...619L..79T} find and discuss sites of recent star formation in the extreme outer disk of M83, associated with the warped \\HI\\ disk of M83, which raises questions about star formation efficiency and modes of star formation. The nature of this outer disk is investigated further in \\citet{2007ApJ...661..115G}, who find evidence that individual young stars are responsible for the UV emission in the outer regions. \\citet{2007AJ....134..135Z} find that these UV sources are quite common out to $2 \\times R_{25}$. In this paper we include candidate PDRs outside the main optical disk of M83 in an effort to probe the amount of available gas in this environment. The suspected high densities of molecular gas in M83, the morphology of the gas in the spiral arms, and the evidence of recent star formation in the outer regions, make this galaxy an excellent target for a study of photodissociated atomic hydrogen and the volume densities of M83's GMCs using our method. In \\S \\ref{sec:method}, we explain our approach and the data we used. In \\S \\ref{sec:results} we present our results. We discuss and summarize these results in \\S \\ref{sec:discussionconlusions}. ", "conclusions": "\\label{sec:discussionconlusions} In this paragraph we compare our results to \\Htwo\\ densities based on CO measurements. Then we discuss background issues, extinction and dust issues. We end with a summary of our findings. \\subsection{CO results comparison} \\label{sub:CO} The study of \\citet{1999ApJ...513..720R} is detailed enough to find typical sizes and masses of GMCs in the eastern arm of M83. These results can be compared to ours with some additional assumptions. Using a CO(1-0) map with kinematic information, they find CO masses and virial masses onindicatingindicating the order of $1 \\times 10^6 M_{\\odot}$. For example, their source 9 (their Table 4) yields a mass derived from CO emission of 7.3 $\\times 10^6 M_\\odot$ and 3.6 $\\times 10^6 M_\\odot$ for the virial masses, at a 50-80\\% uncertainty. This source corresponds to our FUV source no. 13. Since our method does not yield any information on cloud sizes (we only observe HI on the surface of GMCs), nor on kinetic temperatures in these clouds, we will take typical values of our measured volume densities and compare densities using a typical GMC radius of 75 pc for a spherical cloud. $n = 2 n_{H_2}$ inside the GMC and a typical density near our source no. 13 is $20~\\rm{cm}^{-3}$, or $n_{H_2} = 10~\\rm{cm^{-3}}$. 7.3 $\\times 10^6 M_\\odot$ is equivalent to about $n_{H_2} = 85 \\rm{cm^{-3}}$, which is in reasonable agreement with our results considering the uncertainties in both results (70-80\\%). More recent unpublished data (Lord, private communication, 2008) may be quantitatively different and more detailed, but does not yield substantially different results at the resolution of the data that were used in this paper. One of the key differences between M81 and M83 is the abundance of CO emission. While CO is extremely faint and hard to detect in M81 \\citep[e.g.][]{2007A&A...473..771C}, M83 displays bright CO features. The method we used here to find hydrogen densities (and molecular hydrogen densities in GMCs) does not yield any morphological information, since we are observing \\HI\\ on the surface of PDRs. We can note, however, that the size and scale we assume for our candidate PDRs (the range of values of $\\rho_{\\HI}$) is consistent with findings using CO. Since we did not find suitable candidate PDRs in the inner kiloparsec of M83, we cannot compare our results with for example the results of \\citet{1993ApJ...414...98W}, although the densities we find match their low-density component. As to the truncation of the molecular disk that was mentioned by \\citet{2002AJ....123.1892C}, we do see a difference between the inner and outer disk of M83, but it does not seem to be that significant in our results. No differences are seen in candidate PDRs in arm or inter-arm regions, insofar as the large scatter in our results allows us to draw this conclusion. We do not find more than a handful of regions with the densities indicated by \\citet{2007ApJ...664..363H}, although the expected sizes of our candidate PDRs are the same. Finally, the hydrogen densities presented here are similar to the M81 results but extend to higher maximum values. This is consistent with the brighter CO emission in M83, since the higher \\Htwo\\ densities will lead to a greater degree of excitation of the CO molecules. Alternatively, this could also indicate a higher fraction of CO molecules in M83's clouds \\citep[e.g.][]{2007A&A...473..771C}. \\subsection{Background levels, extinction and dust} \\label{sub:FUV} The FUV fluxes seem to be higher in the inner parts of M83 and decreasing (on average) going outward, as is shown in Fig. \\ref{fig:F}. These higher fluxes in the inner parts could be caused by brighter sources, and/or a larger number of sources. Since the individual sources are unresolved, no firm conclusions can be drawn here. Outside $R_{25}$ the measured fluxes are roughly constant. At the same time, the abundances of PDR-produced \\HI\\ surrounding these FUV sources (Fig. \\ref{fig:HI} - note that these are individual measurements, not annular averages) seem to follow a similar trend, consistent with its connection to the photodissociating UV radiation. The FUV source contrast does not vary with galactocentric radius, so while both source and background radiation field decrease in intensity towards the outer regions of M83, their relative strength does not change. Another factor that could influence the FUV radiation is galactic foreground extinction. The higher the applied foreground extinction, the higher the resulting total hydrogen volume density. The 0.52 mag extinction \\citep{1998ApJ...500..525S} towards M83 is significantly less than the 1.37 mag we used towards M81 in \\citet{2008ApJ...673..798H}. The \\citet{1998ApJ...500..525S} extinction correction for M81 would be 0.58, significantly lowering the total hydrogen volume densities. We find a \\correction{wider} range of values in M83, including higher gas densities. We therefore expect M83 to harbor more gas than M81. The M81 and M83 results are hard to compare directly because the extinction corrections were based on different sources in the literature. We intend to compare the two galaxies, together with similar M33 results, more quantitatively in a future paper, with consistent extinction corrections. Our results are also sensitive to the dust-to-gas ratio. The slope of M83's metallicity is relatively shallow. Our preferred dust model assumes that metallicities in the outer parts of M83 remain high. Lower metallicities would result in higher gas densities. Our results in these regions are therefore most likely lower limits and much more gas could be present. The recent results by \\citet{2007ApJ...661..115G} are ambiguous in the sense that the authors provide high metallicity and low metallicity results, depending on the adopted model. In the high metallicity scenario, the slope of M83's metallicity remains shallow out to large galactocentric radius. The low metallicity scenario is accompanied by a sharp drop in metallicity starting at approximately $R_{25}$. \\citet{2000ApJ...538..608S} found a similar range of gas densities in M101 using basically the same method with additional extinction corrections. Their candidate PDRs were all within M101's $R_{25}$, but the FUV luminosities within that range are similar. As in our previous M81 results, no internal extinction correction was applied. If any such correction were applied, it should scale with the dust-to-gas ratio. This means that the extinction would decrease going outward. The FUV fluxes would be even higher in the center of M83 (leading to higher total hydrogen volume densities), while the fluxes in the outer regions of M83 would not be affected much. \\subsection{Summary} \\label{sub:summary} In summary, we have investigated atomic hydrogen found in candidate PDRs in M83 and used the physics of these PDRs to derive total hydrogen volume densities. We carefully considered the contribution of observational selection effects to our results. We find a range of densities: 0.1 - 400 cm$^{-3}$ inside $R_{25}$, 0.5 - 50 cm$^{-3}$ outside $R_{25}$, based on measurements of \\HI\\ believed to be produced in large scale PDRs in M83. The higher GMC volume densities which we find within $R_{25}$ correlate well with the presence of bright CO(1-0) emission there. This points to enhanced collisional excitation as one reason for the CO emission in these GMCs. We note that this is also consistent with our results in M81, where we find little evidence for high GMC volume densities, and for which the CO(1-0) emission is faint. Our measurements go out to a galactocentric radius of 13 kpc (deprojected). Our study used the tilted ring model from \\citet{1974ApJ...193..309R} to get proper galactocentric radii outside the M83 optical disk. Our results are notably sensitive to the local dust-to-gas ratio, especially since the metallicity in the outer regions remains uncertain." }, "0808/0808.3385_arXiv.txt": { "abstract": "% RCW~38 is a uniquely young ($<$1 Myr), embedded ($A_V \\sim 10$) stellar cluster surrounding a pair of early O stars ($\\sim$O5.5) and is one of the few regions within 2~kpc other than Orion to contain over 1000 members. X-ray and deep near-infrared observations reveal a dense cluster with over 200 X-ray sources and 400 infrared sources embedded in a diffuse hot plasma within a 1~pc diameter. The central O star has evacuated its immediate surroundings of dust, creating a wind bubble $\\sim$0.1 pc in radius that is confined by the surrounding molecular cloud, as traced by millimeter continuum and molecular line emission. The interface between the bubble and cloud is a region of warm dust and ionized gas, which shows evidence for ongoing star formation. Extended warm dust is found throughout a 2--3 pc region and coincides with extended X-ray plasma. This is evidence that the influence of the massive stars reaches beyond the confines of the O star bubble. RCW~38 appears similar in structure to RCW~49 and M~20 but is at an earlier evolutionary phase. RCW~38 appears to be a blister compact H{\\small II} region lying just inside the edge of a giant molecular cloud. ", "introduction": "% The evolution of high mass clustered star forming regions is complex and poorly understood. Only the nearby ($\\sim$400 pc), optically revealed, Orion Nebula Cluster (ONC) is well studied (see Muench et al.\\ and O'Dell et al.\\ in this Handbook). Yet, a wide variety of high mass embedded clusters is found within 2 kpc of the Sun. Within this limit, the young cluster RCW~38 (08$^h$59$^m$47.2$^s$ -47$^\\circ$31$'$57$''$ (J2000), {\\it l,b} = $268.03^\\circ,-0.98^\\circ$) is one of the few regions other than the ONC to contain over 1000 members (Lada \\& Lada 2003; Wolk et al.\\ 2006). RCW~38 has an embedded and dense stellar population, comparable to other regions that have been studied recently with Spitzer (e.g., M~20 and RCW~49; Rho et al.\\ 2004, Whitney et al.\\ 2004, Churchwell et al.\\ 2004). RCW~38 provides a unique opportunity to study the evolution of a rich cluster during the phase where its most massive members, a pair of O5.5 stars (DeRose et al.\\ 2008) have just completed their ultracompact {H}{\\small II}\\ region (UC{H}{\\small II}) phase and are now greatly influencing its natal environment and the evolution of its low mass members. ", "conclusions": "RCW~38 appears to be a blister compact HII region with winds originating from the central O-star IRS~2. These winds are excavating the mass in the immediate vicinity of IRS~2 creating a shell--like structure, detected as a radio continuum ring and also evident at infrared and millimeter wavelengths. A few hundred young low mass stars are found in the immediate vicinty of IRS~2, and may be directly exposed to its winds and ionizing radiation. The region as a whole is estimated to contain around 2000 young stars, with 30 OB star candidates. The region of the ring to the west of IRS~2, containing IRS~1, appears to be particularly active and is likely the site of ongoing (triggered) star formation. The IRS~1 ridge appears to be the interface between the IRS~2 wind and either a similar wind from a high mass star further to the west, or a dense clump of gas that is being ablated by IRS~2. Further observations are needed to test these scenarios. In many ways, RCW~38 appears to be a younger more embedded version of the ONC and deserves further study at higher angular resolution and across the spectrum." }, "0808/0808.1349_arXiv.txt": { "abstract": "After the discovery that supermassive black holes (SMBHs) are ubiquitous at the center of stellar spheroids and that their mass $\\Mbh$, in the range $10^6\\Msun -10^9\\Msun$, is tightly related to global properties of the host stellar system, the idea of the co-evolution of elliptical galaxies and of their SMBHs has become a central topic of modern astrophysics. Here, I summarize some consequences that can be derived from the galaxy scaling laws and present a coherent scenario for the formation and evolution of elliptical galaxies and their central SMBHs, focusing in particular on the establishment and maintenance of their scaling laws. In particular, after a first observationally based part, the discussion focuses on the physical interpretation of the Fundamental Plane. Then, two important processes in principle able to destroy the galaxy and SMBH scaling laws, namely galaxy merging and cooling flows, are analyzed. Arguments supporting the necessity to clearly distinguish between the origin and maintenance of the different Saling Laws, and the unavoidable occurrence of SMBH feedback on the galaxy Interstellar Medium in the late stages of galaxy evolution (when elliptical galaxies are sometimes considered as ``dead, red objects''), are then presented. At the end of the paper I will discuss some implications of the recent discovery of super-dense ellipticals in the distant Universe. In particular, I will argue that, if confirmed, these new observations would lead to the conclusion that at early epochs a relation between the stellar mass of the galaxy and the mass of the central SMBH should hold, consistent with the present day Magorrian relation, while the proportionality coefficient between $\\Mbh$ and the scale of velocity dispersion of the hosting spheroids should be significantly smaller than that at the present epoch. ", "introduction": "The mutual interplay between supermassive black holes (hereafter SMBHs) at the center of stellar spheroids\\footnote{The term {\\it early-type galaxies} is generically used for galaxies belonging to the family of {\\it elliptical} galaxies (Es), $S0$ galaxies, {\\it dwarf ellipticals} (dE), and {\\it dwarf spheroidals} (dSph); the class of {\\it stellar spheroids} is made of early-type galaxies and {\\it bulges} of spiral galaxies. For a detailed account of the observational properties of these classes see, e.g.,~\\cite{ref:bm98,ref:bert00}.} and their host systems is now established beyond any reasonable doubt, as indicated by the remarkable correlations found between host galaxy properties and the masses of their SMBHs (e.g.~\\cite{ref:mag98}-\\cite{ref:nfd06}). More specifically, it is now believed that all early-type galaxies with $M_{\\rm B}\\lsim -18$ mag (\\cite{ref:ferrA06}) host a central SMBH (e.g.~\\cite{ref:kr95}-\\cite{ref:dez01}), whose mass $\\Mbh$ scales {\\it linearly} with the spheroid stellar mass $\\Mstar$; the correlation of $\\Mbh$ with the central stellar velocity dispersion $\\sigz$ of the host galaxy is even tighter. It is then natural to argue (e.g.~\\cite{ref:sr98}-\\cite{ref:crEA06}) that the central SMBHs have played an important role in the processes of galaxy formation and evolution, the imprint of which is represented by the Scaling Laws (hereafter SLs) mentioned above. As an additional supporting argument, several groups have noted the link between the cosmological evolution of QSOs and the formation history of galaxies (e.g.~\\cite{ref:kh00}-\\cite{ref:meEA03}, see also~\\cite{ref:pei95}-\\cite{ref:hco04}). In addition to the SLs of their central SMBHs, early-type galaxies are also known to follow well defined empirical SLs relating their global observational properties, such as total luminosity $L$, effective radius $\\re$, and central velocity dispersion $\\sigz$. Among others we recall the Faber \\& Jackson (\\cite{ref:fj76}, hereafter FJ), the Kormendy (\\cite{ref:korm77}), the Fundamental Plane (\\cite{ref:djd87,ref:dreA87}, hereafter FP), the color-$\\sigz$ (\\cite{ref:ble92}), and the Mg$_2$-$\\sigz$ (e.g.~\\cite{ref:buEA88}-\\cite{ref:ber03c}) relations. Clearly, all together these scaling relations reveal the {\\it remarkable homogeneity} of early-type galaxies, provide invaluable information about their formation and evolution, and set stringent requirements that must be taken into account by any proposed galaxy-SMBH formation scenario. We recall that two alternative scenarios for the formation of Es have been proposed. In the {\\it monolithic collapse} picture, ellipticals are formed at early times by dissipative processes (e.g.,~\\cite{ref:els62}-\\cite{ref:lar75}; see also~\\cite{ref:bin77,ref:ro77}), while in the {\\it hierarchical merging scenario} spheroidal systems are the end--products of several merging processes of smaller galaxies, the last major merger taking place in relatively recent times, i.e. at $z \\lsim 1$ (e.g.~\\cite{ref:too77}-\\cite{ref:coEA00}). Each of the two scenarios scores observational and theoretical successes and drawbacks (e.g.~\\cite{ref:jpo80,ref:ren06}). For example, the merging picture (in its {\\it dry} flavor, i.e. neglecting the role of the dissipative gas), could be supported by some observational data suggesting that a fraction of red galaxies in clusters at intermediate redshift are undergoing merging processes: these galaxies could be the progenitors of present-day early-type galaxies (e.g.~\\cite{ref:vdEA99}). At the same time, it is not clear how repeated merging events can produce a class of objects following striking scaling laws involving their global structure, dynamics, and stellar population properties, while theoretical investigations showed that the dynamical processes expected to follow the strongly dissipative phases of monolithic collapse apparently lead to systems surprisingly similar to real Es (e.g., see~\\cite{ ref:va82,ref:bs84}). In this paper I will summarize some selected topics of research in this field and present a possible coherent scenario for the co-evolution of SMBHs and their host spheroids. Because of the enormous body of dedicated literature now available, I will focus my attention on a few selected topics. Very important arguments, such as the dynamics of binary SMBHs and their effects on the central regions of galaxies (e.g.~\\cite{ref:mer06a}-\\cite{ref:mer07} and references therein), or the growth of SMBHs via accretion of stars (e.g.~\\cite{ref:cb94} and references therein) are just mentioned here but not discussed. The paper is organized as follows. After a first, observationally based part (Sect.~2), where the main SLs followed by early-type galaxies and by their central SMBHs are described, in Sect.~3 I will focus on the physical interpretation of the FP: this phenomenological part is a prerequisite to the construction of a possible formation scenario. Then (Sect.~4) I discuss two important physical mechanisms in principle able to destroy the galaxy and SMBH SLs, namely {\\it galaxy merging} and {\\it cooling flows}. In this context, I present some arguments that require that we distinguish between the {\\it origin} and {\\it maintenance} of the SLs and the unavoidable occurrence of SMBH {\\it feedback} on the galaxy Interstellar Medium in the late stages (when Es are sometimes considered as ``dead, red objects''). This last point is strictly related to the solution of the long-standing problem of cooling flows in Es (and clusters), and to the quiescence of Active Galactic Nuclei (AGN) in low-redshift systems. Section 5 addresses the problem of the {\\it origin} of the SLs, while Sect.~6 provides a summary of the main results, with a short discussion of some very recent observational findings, i.e. the surprisingly high stellar density of Es at high redshift. ", "conclusions": "We are finally in the position to connect the different pieces of information described in the previous Sections, to see if it is possible to form a plausible scenario in which the existence of the galaxy and central black holes Scaling Laws described in Sect. 2 can be traced back to the process of galaxy formation. The {\\it first} important clue is that stellar spheroids are a remarkably regular family of stellar systems: the regularity is apparent in terms not only of density profiles, but also of orbit composition and of stellar populations. All these indications point towards a {\\it common} formation mechanism, where the galaxy mass has been a {\\it major} parameter, because many galaxy Scaling Laws involve the galaxy luminosity. In Sect.~3 a review of the different proposed interpretations of the galaxy Scaling Laws has been presented. While a definite answer is not reached yet, it is generally acknowledged that the galaxy SLs are mainly due to a {\\it systematic variation with the galaxy luminosity} of the dark matter amount and distribution, of the light distribution (the so-called weak homology), and finally of the metallicity of the bulk of the stellar mass. The {\\it second} piece of information about galaxy Scaling Laws is indirect and comes from cosmological simulations: by studying simulations of structure formation on the scale of clusters of galaxies, it is found that well defined Scaling Laws are naturally (i.e., by initial conditions) imprinted in the resulting dark matter halos. Thus, it is tempting to put the two points above together and to suggest that the formation of stellar spheroids proceeded mainly in a way similar to the {\\it monolithic collapse} scenario. This would explain the galaxy Scaling Laws just as the imprint of the dark matter halos Scaling Laws (at the mass scale of galaxies) on the baryons. Numerical simulations of fast (dissipationless) collapse in pre-existing dark matter halos can reproduce Sersic profiles similar to those observed, from the outer parts of the models down to their central regions. Cold dissipationless collapse is a process which is expected to dominate the late stages of an initially dissipative process, in which the gas cooling time of the forming galaxy is shorter than its dynamical time, so that stars form ``in flight'', and the subsequent dynamical evolution is dissipationless. Observational evidence supports this argument. In fact, the observed color-magnitude and Mg$_2$-$\\sigz$ relations, and the increase of the $[\\alpha/{\\rm Fe}]$ ratio with $\\sigz$ in the stellar population of Es (e.g.~\\cite{ref:ber03c},\\ref{ref:tgb99}-\\cite{ref:saEA00}), suggest that star formation in massive ellipticals was not only more efficient than in low-mass galaxies, but also that it was faster (i.e., completed before SNIa explosions took place), with the time-scales of gas consumption and ejection shorter or comparable to the galaxy dynamical time (e.g.~\\cite{ref:matt94, ref:pmat04}) and decreasing with increasing galaxy mass. The {\\it third} piece of information is related to the effects of {\\it dry and wet merging} on the Scaling Laws: in fact we know that ellipticals cannot be originated by parabolic merging of low mass spheroids only, even in the presence of substantial gas dissipation (which, at variance with dry merging, is able to increase the galaxy central velocity dispersion, see also \\cite{ref:bnj07})). However, it is also known that SLs such as the FJ, Kormendy, FP, and the $\\Mbh$-$\\sigz$ relations, when considered over the whole mass range spanned by ellipticals in the local Universe, are robust against merging (see also~\\cite{ref:peng07}). Thus the galaxy Scaling Laws, possibly established at high redshift by the fast collapse in pre-existing dark matter halos of gas rich and clumpy stellar distributions (e.g.~\\cite{ref:mgsc07}), can persist even in the presence of a (small) number of dry mergers at lower redshift (\\cite{ref:mhb07}). If this is the case, then monolithic-like collapse at early times and subsequent merging could just represent the different phases of galaxy formation (collapse) and evolution (merging, in addition to the aging of the stellar populations and related phenomena). The possibility that monolithic collapse and successive merging events are just the leading physical processes at different times in galaxy evolution, and that they are both important for galaxy formation, is perhaps indicated also by a ``contradictory'' and often overlooked peculiarity of massive ellipticals. In fact, while the Kormendy relation dictates that the mean stellar density of galaxies decreases with increasing galaxy mass (a natural result of parabolic dry merging), the normalized light profiles of Es becomes steeper and their metallicity increases at increasing galaxy mass (as expected in case of significant gas dissipation). Thus, the present-day light profiles of ellipticals could represent the fossil evidence of the impact of both processes. If the above scenario is correct, then one expects that the star formation history in the Universe and the QSO activity should be roughly parallel. It remains to be clarified if QSO activity brought the process of galaxy formation to an end, or the star formation feedback by ejecting the remaining gas from the galaxy brought the epoch of vigorous QSO activity to a conclusion, or finally if a combination of stellar and AGN feedback was the key factor. My personal view is that we currently have more indications supporting the idea that galaxy formation was stopped more by stellar feedback (i.e., supernova heating) than by the AGN feedback (but see~\\cite{ref:scha07}). In any case, after the end of the fast star formation epoch, {\\it necessarily} a new evolutionary phase begins for the galaxies and their SMBHs. This obvious fact is curiously neglected quite often in the current literature, but it is unavoidable. In fact, over a cosmological time in a passively evolving galaxy the {\\it stellar mass losses} amount, over a cosmological time, to a considerable fraction ($\\gsim 30\\,\\%$) of the total stellar mass, i.e., $\\sim$ 2 orders of magnitude larger than the observed SMBHs masses. If only a minor fraction of this recycled gas (the basic ingredient of the galaxy ``cooling flow'' model!) were accreted on the central SMBH, the SLs involving the BH masses (such as the Magorrian and the $\\Mbh$-$\\sigz$) would be completely different. Thus, the need of an {\\it extremely efficient} feedback from SMBHs is {\\it not} required by complex physical arguments, but just by the {\\it mass budget} of the SMBHs. In addition, this feedback {\\it must} be active over the whole galaxy life, and cannot be temporally concentrated just at the end of the star formation epoch. {\\it However, a moderate accretion from the recycled gas by the evolving stars will not destroy an already established SL, such as the Magorrian relation, as the available ``fuel'' for accretion is naturally proportional to the stellar mass of the host system.} I conclude this Review with a brief comment on a recent and very interesting observational finding, i.e. the fact that apparently stellar spheroids were much {\\it denser} than today at redshift $1\\lsim z\\lsim 2$ (e.g.~\\cite{ref:dise05,ref:truj07,ref:cima08}, and references therein). The obvious question is what mechanism could make a galaxy ``expand''. Of course, internal dynamical processes cannot be invoked, as their time-scales are measured {\\it either} by the galaxy dynamical time (very short compared to the age of the system), {\\it or} by the 2-body relaxation time (which is orders of magnitude longer than the age of the Universe). Thus, the only obvious possibility is to postulate that a few events of dry merging are common in the life of early-type galaxies. This would also help to explain the ``central density-slope paradox'' discussed above. In addition, if dry merging (through a small number of events) is the solution to the problem of superdense galaxies then, in order for present-day galaxies to obey the Magorrian relation, the SMBHs at the center of the superdense progenitors should also follow the same SL, because no significant amount of gas can be accreted on the center in a dry merging. Then, the superdense galaxies cannot follow the $\\Mbh$-$\\sigz$ relation observed in the local Universe because their velocity dispersion is higher than in local galaxies of the same mass. In practice, if superdense galaxies are the progenitors of the nearby Es, and if their expansion was caused by dry merging, they should obey the Magorrian relation observed in the local Universe, but they should fail the local $\\Mbh$-$\\sigz$ relation, by exhibiting a systematically lower zero-point. Testing observationally this conjecture would be an interesting goal for the future." }, "0808/0808.1441_arXiv.txt": { "abstract": "Several short-lived radionuclides (SLRs) were present in the early solar system, some of which should have formed just prior to or soon after the solar system formation. Stellar nucleosynthesis has been proposed as the mechanism for production of SLRs in the solar system, but no appropriate stellar source has been found to explain the abundances of all solar system SLRs. In this study, we propose a faint supernova with mixing and fallback as a stellar source of SLRs with mean lives of $<$5 Myr ($^{26}$Al, $^{41}$Ca, $^{53}$Mn, and $^{60}$Fe) in the solar system. In such a supernova, the inner region of the exploding star experiences mixing, a small fraction of mixed materials is ejected, and the rest undergoes fallback onto the core. The modeled SLR abundances agree well with their solar system abundances if mixing-fallback occurs within the C/O-burning layer. In some cases, the initial solar system abundances of the SLRs can be reproduced within a factor of 2. The dilution factor of supernova ejecta to the solar system materials is $\\sim$10$^{-4}$ and the time interval between the supernova explosion and the formation of oldest solid materials in the solar system is $\\sim$1 Myr. If the dilution occurred due to spherically symmetric expansion, a faint supernova should have occurred nearby the solar system forming region in a star cluster. ", "introduction": "The former presence of short-lived radionuclides (SLRs) in the early solar system ($^{10}$Be, $^{26}$Al, $^{36}$Cl, $^{41}$Ca, $^{53}$Mn, $^{60}$Fe, $^{107}$Pd, $^{129}$I, and $^{182}$Hf) has been inferred from excesses in the abundances of their daughter nuclides in meteorites, linearly correlated with the abundance of a parent element \\citep[e.g.,][]{McD03,Ki05}. These now-extinct SLRs may provide high-resolution ($<$0.1 million years (Myr)) chronometers for events that occurred during the first several million years of solar system evolution and may also be potential heat sources for asteroidal metamorphism and/or differentiation. The SLRs with relatively long mean-lives such as $^{107}$Pd, $^{129}$I, $^{182}$Hf, and perhaps $^{53}$Mn may have been products of steady-state nucleosynthesis in the galaxy \\citep{Ja05}, while those with mean-lives ($\\tau$) of $<$5 Myr, $^{10}$Be ($\\tau$=2.2 Myr), $^{26}$Al ($\\tau$=1.03 Myr), $^{36}$Cl ($\\tau$=0.43 Myr), $^{41}$Ca ($\\tau$=0.15 Myr), $^{60}$Fe ($\\tau$=2.2 Myr), and possibly $^{53}$Mn ($\\tau$=5.3 Myr), should have been produced either by energetic-particle irradiation in the early solar system or by stellar nucleosynthesis just prior to or shortly after the birth of the solar system. Beryllium-10, which was found in CAIs (calcium-aluminum-rich inclusions) (McKeegan et al., 2000), is not produced by stellar nucleosynthesis, but can be efficiently formed by energetic particle irradiation. On the other hand, $^{60}$Fe, the abundance of which in the early solar system also appears to require a late addition \\citep{TH03,Mo05,Ta06}, can be efficiently produced only in stars. Thus, the presence of $^{10}$Be and $^{60}$Fe in the early solar system suggests that both energetic particle irradiation and stellar nucleosynthesis make contributions to the inventories of solar system SLRs. However, it is not clear yet which process contributed more significantly to the inventory of the SLRs that could be synthesized by both processes, such as $^{26}$Al, $^{36}$Cl, $^{41}$Ca, and $^{53}$Mn. There have been several attempts to find a plausible stellar source(s) for the abundances of SLRs in the early solar system. A low-mass (1.5 M$_{\\sun}$) thermally-pulsing asymptotic-giant-branch (TP-AGB) star cannot produce enough $^{60}$Fe to match the initial abundance of $^{60}$Fe in the solar system \\citep[e.g.,][]{Bu03,Wa06,SS06}. Models for intermediate-mass AGB stars (5 M$_{\\sun}$ with the solar metallicity or 3 M$_{\\sun}$ with 1/3 $\\times$ the solar metallicity) could explain the inferred solar system abundance of $^{60}$Fe as well as the abundances of $^{26}$Al, $^{41}$Ca and $^{107}$Pd \\citep{Wa06}. However, the probability of encounters between molecular clouds and AGB stars seems to be extremely low \\citep{KM94}, implying that an AGB star was an unlikely source of SLRs in the solar system. Mass-loss winds from Wolf-Rayet stars may have been a source of the solar system $^{26}$Al, $^{36}$Cl, $^{41}$Ca, and $^{107}$Pd if the nuclides were incorporated into the solar system within a time-interval of $\\sim$10$^{5}$ year after production \\citep{Ar06}. However, a WR wind alone would not contain enough $^{60}$Fe and $^{53}$Mn to explain their solar system initial abundances \\citep{Ar06}. Type II core-collapse supernovae have also been considered as plausible sources for SLRs. However, most supernova models imply that if a supernova provided $^{26}$Al and $^{41}$Ca into the solar system, it would also supply 10-100 times more $^{53}$Mn than its estimated initial abundance in the solar system \\citep[e.g.,][]{GV00,Ou05,SS06}. This discrepancy could be explained by fallback of the innermost layers containing most of the $^{53}$Mn onto a collapsing stellar core \\citep{MC00,Me05}. In such a case, $^{53}$Mn could primarily be derived from a different source such as the interstellar medium. \\cite{Wa06} proposed that $^{53}$Mn and possibly $^{60}$Fe were injected into the solar system as supernova ejecta with a time-interval of $\\sim$10$^{7}$ years after production, long enough for $^{41}$Ca and $^{26}$Al to decay completely, and that $^{26}$Al and $^{41}$Ca in the solar system may have been produced either by energetic particle irradiation or by an AGB star. Another problem with supernovae as sources of SLRs is overproduction of $^{60}$Fe if all the $^{26}$Al in the solar system was derived from supernovae. Although the yield of $^{60}$Fe depends the mass loss and initial mass \\citep[e.g.,][]{LC06}, the expected amount of $^{60}$Fe injected from a supernova would be, in general, a few times higher than its highest estimate for the early solar system, as we will show later. This problem may still remain even in models for fallback supernovae. In this study, we propose a supernova with mixing and fallback, with a kinetic energy of explosion slightly less than that for a typical supernova ($\\sim$10$^{51}$ erg) as a potential source of $^{26}$Al, $^{41}$Ca, $^{53}$Mn, and $^{60}$Fe in the early solar system. Faint supernovae such as SN1997D and SN1999br have such kinetic energies \\citep[e.g.,][]{No06}. In models for supernovae with mixing-fallback, the inner region of the exploding star experiences mixing, some fraction of mixed materials is ejected, and the rest undergoes fallback onto the core \\citep[e.g.,][]{UN02,UN05,No06,To07}. Nucleosynthesis in a faint supernova with mixing-fallback successfully reproduces the elemental abundance patterns of hyper metal-poor stars \\citep{UN03,Iw05,No06}. ", "conclusions": "" }, "0808/0808.3358_arXiv.txt": { "abstract": "{We present large scale 870\\,\\um\\ maps { of} the nearby starburst galaxies NGC\\,253 and NGC\\,4945 as well as the nearest giant elliptical radio galaxy Centaurus A (NGC\\,5128) obtained with the newly commissioned {Large Apex Bolometer Camera (LABOCA)} operated at the { Atacama Pathfinder Experiment} telescope. Our continuum images reveal for the first time the distribution of cold dust at a angular resolution of { 20$''$} across the entire optical disks of NGC\\,253 and NGC\\,4945 out to a radial distance of $10'$ (7.5 kpc). In NGC\\,5128 our LABOCA image also shows, for the first time { at submillimeter wavelengths}, the synchrotron emission associated with the radio jet and the inner radio lobes. From an analysis of the 870\\,\\um\\ emission in conjunction with ISO-LWS, IRAS and long wavelengths radio data we find temperatures for the cold dust in the disks of all three galaxies of 17--20\\,K, comparable to the dust temperatures in the disk of the Milky Way. The total gas mass in the three galaxies is determined to be 2.1, 4.2 and $2.8\\times10^{9}\\,\\msol$ for NGC\\,253, NGC\\,4945 and NGC\\,5128, respectively. The mass of the warmer (30--40\\,K) gas associated with the central starburst regions in NGC\\,253 and NGC\\,4945 only accounts for $\\sim10\\%$ of the total gas mass. A detailed comparison between the gas masses derived from the dust continuum and the integrated CO(1--0) intensity in NGC\\,253 suggests that changes of the CO luminosity to molecular mass conversion factor are mainly driven by a metallicity gradient and only to a lesser degree by variations of the CO excitation. An analysis of the synchrotron spectrum in the northern radio lobe of NGC\\,5128 shows that the synchrotron emission from radio to the ultraviolet (UV) wavelengths is well described by a broken power law and that the break frequency is a function of the distance from the radio core as expected for aging electrons. We derive an outflow speed of $\\sim0.5$\\,{\\it c} at a distance of 2.6\\,kpc from the center, consistent with the speed derived in the vicinity of the nucleus. ", "introduction": "Nearly half the bolometric luminosity in the local universe is emitted at mid- and far-infrared (IR) wavelengths. The IR radiation is produced by warm interstellar dust grains that are heated by UV photons from hot high mass stars. This thermal emission of dust grains therefore carries valuable information on feedback processes from star formation, the chemical composition of the { interstellar medium (ISM)} and also on the total amount of interstellar matter in galaxies, the gas surface density and its relation to star-forming regions. The high spatial resolution of the Spitzer Space Telescope has greatly improved our view in this important { wavelength} range and allows for the first time to study spatially resolved spectral energy distributions (SEDs) across galaxies and { the relation of IR emission} to other tracers of star formation. \\\\ In order to provide a complete measurement of the dust SED it is highly desirable to combine the IR observations with data on the long wavelength (Rayleigh-Jeans) tail of the SED in the submm regime (see, e.g., Draine \\etal\\ \\cite{draine07}). This is because the submm observations are particularly sensitive to cold dust which dominates the dust mass in galaxies. \\\\ Ground based observations of the submm emission of nearby galaxies, however, remain a challenging task because they largely suffer from limitations due to the Earth's atmosphere. Furthermore, existing submm cameras have been limited in their field of view (FoV) to a few arc minutes which makes it difficult to survey large areas on the sky in a reasonable amount of time.\\\\ With the commissioning of the Large APEX Bolometer Camera (LABOCA, Siringo \\etal\\ \\cite{siringo07} and in prep.) at the APEX telescope (G\\\"usten \\etal\\ \\cite{guesten06}) at 5100\\,m altitude on Chajnantor this situation has largely improved. With its large field of view and large number of detectors in combination with the extremely dry conditions at the site, LABOCA provides the mapping-speed and sensitivity required to survey large areas on the sky down to mJy noise levels.\\\\ In this paper we present the first large scale 870\\,\\um\\ maps { by LABOCA} towards two nearby starburst galaxies NGC\\,253 and NGC\\,4945, and towards the nearest giant elliptical radio galaxy Cen\\,A (NGC\\,5128). In Sect.\\,2 we describe the LABOCA observations and the data reduction, Sect.\\,3 focuses on the dust SEDs and gas masses. In Sect.\\,4 we discuss the thermal and non-thermal emission processes in our target galaxies and Sect.\\,5 summarizes our results. ", "conclusions": "\\subsection{Dust temperatures} Our dust temperature estimate for the coldest, mass carrying dust component of 17\\,K in the disk of NGC\\,253 is in excellent agreement with estimates based on lower resolution ISOPHOT and IRAS data (Melo \\etal\\ \\cite{melo02}, Radovich \\etal\\ \\cite{radovich01}). The dust SEDs in the disks of NGC\\,4945 as well as NGC\\,5128 show similarly low dust temperatures, which suggests that the temperatures of the dust in the disks of all three active galaxies are comparable to those found in the Milky Way (MW, e.g. Cox \\& Mezger \\cite{cox89}). None of our target shows evidence for dust with temperatures below 10\\,K. This finding is also in agreement with results based on SCUBA and Spitzer observations of the SINGS galaxy sample (Draine \\etal\\ \\cite{draine07}). { Our dust SEDs, however, do not rule out the presence of such cold gas. Its effect on the total dust/gas mass estimate, on the other hand, is not dramatic because a cold (e.g. a 8\\,K) dust component in the disks cannot contribute to more than $\\sim$40\\% of the observed 870\\microns\\ flux in order to be consistent with the overall SED shape.}\\\\ For the nuclear starburst regions our analysis suggests that NGC\\,253 is free of cold (T$<$30\\,K) dust while the central SEDs for NGC\\,4945 and NGC\\,5128 both show a clear signature of dust at temperatures comparable to those found in the disks. This, however, does not necessarily imply that the nuclear starburst itself contains cold gas because the 80$''$ aperture analyzed here is still too large to separate spatially the central active region from the surrounding disk. For NGC\\,5128 the situation is further complicated by the strong non-thermal contribution from the central AGN and { the jets} to the 870\\,\\um\\ flux, which makes it difficult to determine the submm flux related to thermal dust emission alone. Therefore, the suggested decrease of the cold { dust component's temperature} from the disk towards the center in NGC\\,5128 may simply reflect an imperfect separation between the thermal and non-thermal flux contributions. \\subsection{Gas masses} Due to the large areas our target galaxies project on the sky, only few mm/submm observations have been published so far which allow to determine the total molecular gas content across the entire optical disks. { For NGC\\,253, Houghton \\etal\\ (\\cite{houghton97}) have presented CO(1--0) observations having a similar extent} than our LABOCA maps. They find a total molecular gas mass of $2.4\\times 10^{9}\\,\\msol$ using a CO conversion factor of $3\\times10^{20}\\,{\\rm cm}^{-2}\\,({\\rm K\\,\\kms})^{-1}$. This value includes the contribution of heavier elements. The \\hi\\ mass of NGC\\,253 is $1.8\\times 10^{9}\\,\\msol$ (Koribalski \\etal\\ \\cite{kori04}) which yields a total mass of $4.2\\times 10^{9}\\,\\msol$, a factor of 2 higher than our estimate based on the dust continuum. This may argue for a smaller CO conversion factor or a higher gas-to-dust mass ratio. { Several studies have suggested that the average gas-to-dust mass ratio of a galaxy is a function of its enrichment (e.g. Draine \\etal\\ \\cite{draine07}, Engelbracht \\etal\\ \\cite{engel08}). Indeed the characteristic oxygen abundance (a measure for the average metallicity of a galaxy) in NGC\\,253 is lower by $\\sim 0.1$\\,dex than the MW's metallicity (Pilyugin \\etal\\ \\cite{pil04}). But such a small difference in metallicity is not expected to cause the higher gas-to-dust mass ratio in NGC\\,253 illustrated in Fig.\\,\\ref{dust_metall} where we show the location of our target galaxies on the dust-to-gas mass ratio vs. characteristic oxygen abundance plot for the SINGS galaxies (Draine \\etal\\ \\cite{draine07}, their Fig\\,16). NGC\\,253 lies well within the factor 2 scatter band on this metallicity relation observed for the SINGS galaxies. Given the simplifying assumptions made in our mass estimate (e.g. the constant dust absorption coefficient across the disk of NGC\\,253 and the use of constant CO conversion factor), the mass estimates for the dust continuum and the CO and \\hi\\ are in reasonable agreement.}\\\\ \\begin{figure} \\centering \\includegraphics[width=8.5cm,angle=0]{9909fig8.eps} \\caption{Dust-to-gas mass ratio vs. characteristic oxygen abundance plot for the SINGS galaxies adopted from Draine \\etal\\ (\\cite{draine07}, their Fig.\\,16). Our target galaxies are shown as filled squares. The CO intensities have been converted to \\hh\\ masses using a conversion factor of $2.0\\times10^{20}\\,{\\rm cm}^{-2}\\,({\\rm K\\,\\kms})^{-1}$.} \\label{dust_metall} \\end{figure} \\noindent CO(1--0) observations covering the optical disk of NGC\\,4945 have been published by Dahlem \\etal\\ (\\cite{dahlem93}). They give an \\hh\\ mass of $2.2\\times 10^{9}\\,\\msol$ (scaled to D\\,=\\,3.8\\,Mpc and using a CO conversion factor of $2\\times10^{20}\\,{\\rm cm}^{-2}\\, ({\\rm K\\,\\kms})^{-1}$). Correcting for heavier elements and taking the \\hi\\ mass of $1\\times 10^{9}\\,\\msol$ (Ott \\etal\\, \\cite{ott01}, scaled to D\\,=\\,3.8\\,Mpc) into account yields a total gas mass of $4\\times 10^{9}\\,\\msol$, in good agreement with our results. { Our dust mass is also in agreement with the expected metallicity dependence for MW like dust (see Fig\\,\\ref{dust_metall}). For the determination of the characteristic oxygen abundance we have used here the O/H --$M_{B}$ relation for spiral galaxies from Pilyugin \\etal\\ (\\cite{pil04}).}\\\\ \\\\ For NGC\\,5128 the situation is different as our estimated gas mass based on the dust emission exceeds those based on CO and \\hi\\ by more than a factor of 4: Eckhart \\etal\\ (\\cite{eckhart90}) derive a total molecular gas mass based on CO of $2.7\\times 10^{8}\\,\\msol$ (scaled to D\\,=\\,3.5\\,Mpc). The \\hi\\ mass accounts for $2.5\\times 10^{8}\\,\\msol$ (Gardner \\& Whiteoak \\cite{gardner76}, scaled to D\\,=\\,3.5\\,Mpc). Correcting for heavier elements this yields a total gas mass of $6\\times 10^{8}\\,\\msol$ compared to our estimate of $2.8\\times 10^{9}\\,\\msol$. We note that this discrepancy still holds if we would assume that the cold dust towards the center has a temperature of 20\\,K instead of the 14\\,K derived from our SED fitting (see discussion above).\\\\ Based on SCUBA 850 and 450\\,\\um\\ observations, Leeuw \\etal\\ (\\cite{leeuw02}) find a dust mass of $2.2\\times 10^{6}\\,\\msol$ for the central 4.5\\,kpc which translates into a gas mass of $3.3\\times 10^{8}\\,\\msol$ assuming a gas-to-dust mass ratio of 150 -- a factor of two lower than the estimate based on CO and a factor of 8 lower than our estimate. This discrepancy is difficult to understand because our assumed dust absorption coefficient, $\\kappa_\\nu$, differs at most by $\\sim30\\%$ from the values used by Leeuw \\etal\\ and our LABOCA flux is only slightly higher than the SCUBA 850\\,\\um\\ flux (7.8 Jy compared to 6.4 Jy adding the inner and outer disk as defined in Leeuw \\etal). Furthermore, our dust temperature is somewhat higher than the 12\\,K derived by Leeuw \\etal, so that we would expect only a small difference between the two dust mass estimates. Using the dust temperatures, emissivities and 850\\,\\um\\ fluxes given in Leeuw \\etal\\ we indeed estimate dust and gas masses of $1.1\\times 10^{7}\\,\\msol$ and $1.7\\times 10^{9}\\,\\msol$ for NGC\\,5128 in better agreement with our results. Our higher dust masses are also in agreement with the results of Mirabel \\etal\\ (\\cite{mirabel99}). Their 850\\,\\um\\ fluxes, however, are about a factor of 10 higher than those published by Leeuw \\etal\\ (\\cite{leeuw02}) and from our LABOCA maps and their dust mass should therefore be treated with some caution. Independent estimates of the dust mass in NGC\\,5128 have been derived by Block \\& Sauvage (\\cite{block00}) using the V-15\\,\\um\\ ratio which results in $M_{\\rm d}=2.3\\times10^{6}\\,\\msol$ (scaled to D\\,=\\,3.5\\,Mpc). This study, however, only addresses the dust content of the central 90$''$ and may miss diffuse dust which could significantly increase the estimated dust mass. \\\\ Considering these uncertainties in the dust mass in Cen\\,A it is impossible to obtain strong conclusions by comparing the gas mass derived from CO and the dust continuum. We note, however, that the molecular gas mass derived by Eckart \\etal\\ (\\cite{eckhart90}) is based on a standard galactic conversion factor and therefore unlikely to underestimate the gas mass by more than a factor of 2. Therefore, the high dust mass derived from our LABOCA data may suggest that the dust properties or the gas-to-dust mass ratio in NGC\\,5128 are different from those in NGC\\,253 and NGC\\,4945.\\\\ { To our knowledge no oxygen abundance estimates across the disk of NGC\\,5128 has been published which would allow secure derivation of the characteristic oxygen abundance to check for metallicity effects as a possible reason for the low gas-to-dust mass ratio. Estimates based on the globular cluster systems of NGC5128 indicate metalicities similar to or somewhat higher than their MW counterparts (see Israel \\etal\\ \\cite{israel98} and references therein). Using, as for NGC\\,4945, $M_B$ as indicator for O/H leads to a similar result (12+log(O/H)\\,=\\,--8.5) although it is not clear if this relation holds for elliptical galaxies. We therefore assume a characteristic oxygen abundance similar to the MW which places NGC5128 well above the expected metallicity dependence in Fig.\\,\\ref{dust_metall}.} \\subsection{CO intensity vs dust column density in NGC\\,253} Among our target galaxies the CO emission in NGC\\,253 has been studied in most detail and large scale CO maps exist in the literature with similar spatial resolution than our LABOCA data. Sorai \\etal\\ (\\cite{sorai00}) imaged large parts of the optical disk in CO(1--0) at a spatial resolution of $16''$ using the Nobeyama 45m telescope. { To investigate variations of the mass tracing capabilities of CO in comparison to the dust continuum we have compared their integrated CO intensities to the \\hh\\ column densities derived from our LABOCA maps within the central 6\\,Kpc of NGC\\,253 taking the neutral gas mass fraction measured from \\hi\\ into account. The gas mass (\\hi\\ \\& \\hh) distribution from the dust was computed using a dust temperature distribution as suggested from our dust SED analysis: 35\\,K within the central $30''\\times 16''$, 19\\,K out to a radial distance of $40''$ (ISO aperture) and 17\\,K for emission further out and a constant gas-to-dust mass ratio of 150. This gas mass distribution was corrected for \\hi\\ using the radial dependence of molecular gas mass fraction from Sorai \\etal\\ (\\cite{sorai00}, their Fig.\\,9\\,a) and finally converted to a \\hh\\ column density map.} \\noindent The comparison was done for { 20$''$ sized pixels after smoothing both data sets to that resolution} and is shown in Fig.\\,\\ref{xco}. From this figure there is no obvious difference between the CO conversion factor in the central starburst region and positions in the bar/disk region of NGC\\,253. A linear fit yields conversion { factors of $1.7\\times10^{20}\\,{\\rm cm}^{-2}\\,({\\rm K\\,\\kms})^{-1}$ (2.7\\,\\msol\\,(K\\,\\kms\\,pc$^2$)$^{-1}$) in both regions with errors of 3\\% and 7\\% for the central and the bar/disk region}. { The larger error for the slope outside the nuclear starburst region is mainly due to the smaller dynamic range of the data points. The result is little dependent on the neutral gas correction. For example, if we assume that all gas is in molecular form and ignore the \\hi\\ (which maximizes the conversion factor), we find the same value for the nuclear region and an increase of only 50\\% ($2.5\\times10^{20}\\,{\\rm cm}^{-2}\\,({\\rm K\\,\\kms})^{-1}$) for the bar/disk region.} \\noindent This result is surprising at the first glance, as several detailed CO studies have suggested that the CO conversion factor is { several times} lower in the central starburst region { (e.g. Mauersberger \\etal\\ \\cite{mauers96}, Harrison \\etal\\ \\cite{harrison99} for the center of NGC\\,253)} than the galactic conversion factor, which { in turn} is expected to be a good approximation for the regions outside the starburst. { In contrast Fig.\\,\\ref{xco} hints at a constant conversion factor over the entire galaxy, including its nucleus. One must note, however, that in the strict sense, Fig.~\\ref{xco} implies a constant ratio only for the measured CO/dust masses (or fluxes) across the galaxy, while the conversion of these into total gas mass may both also depend on other characteristics. Specifically, our result is consistent with the possibility of both the gas/CO and the gas/dust ratios varying similarly with metalicities. For example, Mauersberger \\etal\\ (\\cite{mauers96}) assumed a metallicity of 3 times the solar value for their determination of the CO conversion factor in NGC\\,253. Matching this, a gas-to-dust mass ratio that scales inversely with metalicity, i.e.~50 (if a galactic value of 150 is assumed) for the nucleus of NGC\\,253, is implied by our results. The apparent constancy of the CO/dust ratio across NGC253 also suggests that changes in the gas/CO conversion are unlikely to be dominated by the variation of CO excitation, since that would be apparent in deviations from the strict linear relation, which we do not see. } \\begin{figure*} \\centering \\includegraphics[width=16.0cm,angle=0]{9909fig9.eps} \\caption{Observed integrated CO(1--0) intensities (Sorai \\etal\\ \\cite{sorai00}) compared to the \\hh\\ gas column densities derived from our dust continuum analysis in NGC\\,253 { corrected for the \\hi\\ contribution}. Filled squares show positions towards the central starburst region, open circles positions in the bar/disk of NGC\\,253. The dashed line is a linear fit to data in the starburst region only, the dotted line a fit to all other data points. The right part of the figure is a zoom to the full diagram shown on the left.} \\label{xco} \\end{figure*} \\subsection{Radio lobes in Cen\\,A\\label{radiolobes}} The 870\\,\\um\\ emission associated with the radio lobes is very similar in morphology to the 1.4GHz continuum emission (Fig.\\,\\ref{cena_image}\\,right \\& Fig.\\,\\ref{cena_lobes}). As the synchrotron emission from the northern radio lobe has also been detected at 15--4.5\\,\\um\\, in the UV and at X-ray energies (Brookes \\etal\\ \\cite{brookes06}, Hardcastle, Kraft \\& Worrall \\cite{hardcastle06}) it is clear that the 870\\,\\um\\ emission is synchrotron in nature and not associated with dust driven out from the disk by the radio jet. The southern radio lobe is detected for the first time at wavelengths shorter than 3.5\\,cm. We show a high contrast signal to noise map of our LABOCA observations in Fig.\\,\\ref{cena_lobes}. From this image it is evident that the northern radio jet is also detected at 870\\,\\um. The jet, however, is located in a low level negative bowl surrounding the strong central emission. Such negative artifacts are unavoidable byproducts of the skynoise removal which results in the filtering of low spatial frequencies. As a consequence the true 870\\,\\um\\ flux of the jet is difficult to determine. The 870\\,\\um\\ flux of the northern and southern lobes are 2.5 and 0.9\\,Jy respectively.\\\\ The variation of the spectral index of the northern lobe with wavelength has recently been investigated by Hardcastle, Kraft \\& Worrall (\\cite{hardcastle06}) and Brookes \\etal\\ (\\cite{brookes06}). Based on { Spitzer/IRAC}, GALEX and Chandra data Hardcastle \\etal\\ fit the synchrotron emission by a broken power law. Their model with a break frequency at $\\sim 300\\,\\rm{GHz}$, however, underestimates the observed UV flux and also provides a poor fit to the observed slope between the IRAC bands. We have included our LABOCA fluxes in the analysis of the spectral index in the three regions in the northern lobe defined by Hardcastle \\etal\\ (see Fig.\\,\\ref{cena_lobes}). These regions cover the radio jet (inner), the southern part of the radio lobe (middle) and the northern part of the lobe (outer region). The LABOCA flux within the three regions is $40\\pm20$\\,mJy, $365\\pm40$\\,mJy and $1.0\\pm0.1$\\,Jy for the inner, middle and outer region respectively. The synchrotron spectra of the three regions are shown in Fig.\\,\\ref{power}. We fit the radio to X-ray data of all three regions with a non-standard ($\\Delta\\alpha\\not=0.5$), broken power law. For the inner region all data can well be fit with a break frequency of $\\rm{log}_{10}(\\nu_{\\rm b/Hz,inner})=14.5\\pm0.65$ (315\\,THz) and a change of the power index of $\\Delta\\alpha_{\\rm inner}=0.8\\pm0.3$. { Our LABOCA flux falls somewhat below the expected value from this model (68\\,mJy) which we attribute to the difficulties to determine the 870\\,\\microns\\ flux density in this particular region as described above. Because the break frequency in the inner region falls into the NIR regime this data point, however, is not critical for the fitting.} For the middle and the outer regions the break frequency shifts, as expected, to lower frequency yielding $\\rm{log}_{10}(\\nu_{\\rm b/Hz,middle})=12.5\\pm0.3$ (2.9\\,THz) and $\\rm{log}_{10}(\\nu_{\\rm b/Hz,outer})=11.7\\pm0.6$ (500\\,GHz). The change of the power index for both regions is close to that predicted by a continuous injection model (KP model) with $\\Delta\\alpha_{\\rm middle}=0.6\\pm0.1$ and $\\Delta\\alpha_{\\rm outer}=0.6\\pm0.15$. For both regions the broken power law model, however, overpredicts the X-ray flux, which indicates that the energy distribution of the electrons is more complex than a distribution yielding a simple broken power-law. A detailed modeling of the energy distribution is beyond the scope of this paper but we note, that also a single injection (SI) model with its exponential decrease at high energies does not provide a good fit to the data. From Fig.\\,\\ref{power} it can be seen that a model intermediate between the SI and KP, e.g. a broken power law with a high energy exponential decrease, presumably gives a better description of the observed fluxes between the radio and X-ray.\\\\ For the inner jet region, where the broken power law fits all observed fluxes, we can also estimate the speed of the electrons using our fit to the break frequency: assuming a magnetic field strength of 3\\,nT, the minimum energy value (Brookes \\etal\\ \\cite{brookes06}), we get an age of the electrons of $\\sim\\,17000$\\,yrs which yields a speed of 0.5\\,{\\it c} using the measured projected distance of 2.6\\,kpc from the nucleus and a jet view angle of $\\sim20^{\\circ}$ (Hardcastle \\etal\\ \\cite{hardcastle03}). This speed is consistent with the electron speed derived close to the central source. (Hardcastle \\etal\\ \\cite{hardcastle03}). \\begin{figure} \\centering \\includegraphics[width=9.0cm,angle=0]{9909fig10.eps} \\caption{High contrast representation of the signal to noise map of our 870\\,\\um\\ observations towards NGC\\,5128 smoothed to an effective spatial resolution of 25$''$ (beam smoothed). { The negative parts of the image are highlighted by two grey contours with signal to noise ratios of $-1$ and $-2$.} This presentation also shows the faint submm emission from the northern radio jet. The white and black contours are the 1.4\\,GHz VLA map from Condon \\etal\\ (\\cite{condon96}). The { white} boxes indicate those regions which have been used to analyze the spectral index.} \\label{cena_lobes} \\end{figure} \\begin{figure} \\centering \\includegraphics[width=8.5cm,angle=0]{9909fig11.eps} \\caption{Synchrotron spectra from radio to X-rays in three apertures in the northern radio lobe. The apertures are shown in Fig.\\,\\ref{cena_lobes} and correspond to the \"outer\" (shown as dashed line and square symbols), \"middle\" (solid, circles) and \"inner\" (dotted, triangles) region defined in Hardcastle, Kraft \\& Worrall \\cite{hardcastle06}. The filled symbols represent the 870\\,\\um\\ flux densities. For all other data see Table\\,1 in Hardcastle, Kraft \\& Worrall (\\cite{hardcastle06}). The lines show broken power law fits to the data. The presentation for the outer/inner region have been up/down scaled by 1.5 dex.} \\label{power} \\end{figure}" }, "0808/0808.3391_arXiv.txt": { "abstract": "Correlation studies of prompt and afterglow emissions from gamma-ray bursts (GRBs) between different spectral bands has been difficult to do in the past because few bursts had comprehensive and intercomparable afterglow measurements. In this paper\\footnote{to appear in The Astrophysical Journal, Dec 20, 2008, v. 689, no. 2} we present a large and uniform data set for correlation analysis based on bursts detected by the {\\it Swift} mission. For the first time, short and long bursts can be analyzed and compared. It is found for both classes that the optical, X-ray and gamma-ray emissions are linearly correlated, but with a large spread about the correlation line; stronger bursts tend to have brighter afterglows, and bursts with brighter X-ray afterglow tend to have brighter optical afterglow. Short bursts are, on average, weaker in both prompt and afterglow emissions. No short bursts are seen with extremely low optical to X-ray ratio as occurs for ``dark'' long bursts. Although statistics are still poor for short bursts, there is no evidence yet for a subgroup of short bursts with high extinction as there is for long bursts. Long bursts are detected in the dark category at the same fraction as for pre-{\\it Swift} bursts. Interesting cases are discovered of long bursts that are detected in the optical, and yet have low enough optical to X-ray ratio to be classified as dark. For the prompt emission, short and long bursts have different average tracks on flux {\\it vs} fluence plots. In {\\it Swift}, GRB detections tend to be fluence limited for short bursts and flux limited for long events. ", "introduction": "One of the longest enduring Gamma Ray Burst (GRB) classification schemes is based on their distributions in duration and spectral hardness. Both quantities seem to cluster into two separate classes with the longer events (those above $\\sim2$ s; Kouveliotou et al. 1993) being predominantly softer while the shorter ones are harder. The mechanism for the origin of the GRB explosions (the central engine) appears to be quite different for the two types. % Long bursts are ascribed to the core collapse to a black hole of a massive, young, rapidly rotating star in the ``collapsar'' model (Woosley 1993; MacFadyen \\& Woosley 1999; Woosley \\& Bloom 2006) which is supported by observations such as the coincidence of SNe with well-observed nearby GRBs (Galama et al. 1998; Bloom et al. 1999; Staneck et al. 2003; Hjorth et al. 2003; Pian et al. 2006). The prevalent model for short bursts has them caused by the coalescence of a binary pair of compact old stars (Lattimer \\& Schramm 1974; Paczy\\'nski 1986; Eichler et al. 1989; Mochkovitch et al. 1993; Rosswog, Ramirez-Ruiz, \\& Davies 2003; Oechslin, Janka, \\& Marek 2007) which is supported by recent observations of progenitor sites with low star formation activity (Gehrels et al. 2005; Bloom et al. 2006; Fox et al. 2005; Villasenor et al. 2005, Hjorth et al. 2005; Barthelmy et al. 2005b; Berger et al. 2005). In both scenarios, a highly-relativistic collimated outflow of particles and radiation occurs producing prompt gamma-ray emission from shock accelerated electrons, which evolves into a long-lasting afterglow from shock interactions with the circumburst medium (e.g., M\\'esz\\'aros \\& Rees 1997). For short bursts there are also models for the afterglow in which a radioactive wind causes emission in the first day or so (Li \\& Paczy\\'nski 1998, Kulkarni 2005). Correlation studies of prompt and afterglow emission are crucial for understanding their production mechanisms and environmental effects. For example, Jakobsson et al. (2004) developed a criterion for determining which GRBs are ``dark'' bursts, by comparing the relative intensity of their X-ray and optical afterglows to find what fraction of bursts have high column densities. Stratta et al. (2004) studied the X-ray and optical absorption properties of 13 GRBs studied by {\\it BeppoSax}. Roming et al. (2006) and Fynbo et al. (2007) expanded on previous work to include (long) bursts from the {\\it Swift} satellite. A more detailed work on dark bursts using a broad-band spectral analysis is given by Rol et al. (2005, 2007). Zhang et al. (2007) present a study comparing radiative efficiencies for short and long bursts as derived from a correlation analysis. Using {\\it Swift} short bursts, Berger (2007) compared their X-ray afterglow to their gamma-ray prompt emission, and found that 20\\% have anomalously low X-ray to gamma ray ratios indicating very low density burst sites, possibly in globular clusters, for that subpopulation (see also Berger et al 2007). Other correlation studies have been undertaken by Salmonson \\& Galama (2002), Firmani et al. (2006), Nava et al. (2006), Butler (2007), and Nysewander, Fruchter, \\& Pe'er (2008). An early study of X-ray afterglow properties at $t=11$ hr was carried out by Piran et al. (2001). In this study we perform correlation studies using the extensive data set from {\\it Swift}. Sections 2 and 3 cover observations and results, respectively, while in Section 4, we discuss the implications of the results and in Section 5 the conclusions and future prospects. ", "conclusions": "\\subsection{Correlations \\& Short/Long Distributions} We show in this work that correlations exist between prompt and afterglow fluxes of GRBs and between different wavelength bands in the afterglow. The highest significance correlation is between the prompt emission gamma-ray fluence and the X-ray afterglow flux at a significance level of 99.9999996\\% for long bursts and 69\\% for short bursts. The correlation between the optical afterglow and X-ray afterglow fluxes is less significant at 99\\% significance for long bursts and only $\\sim30$\\% for short bursts (for a small sample, however). % It is important to note that there is a wide spread in the data for all of the correlations. The correlations are real and significant, but the fraction of the observed variations due to the correlations between the above parameters accounts for only a portion of the data spread. The correlation can only be used to predict a flux level to within approximately an order of magnitude. The fraction of the variation due to the correlations is given by the square of the correlation parameter, $r$, which, as shown in Table 4, varies from a few percent to 50\\%. The rest of the data spread is due to other factors such as correlations with additional unknown parameters. An example of an additional parameter is extinction in the optical afterglow. Short bursts are weaker on average than long bursts in afterglow fluxes. There is overlap with the dimmer long bursts, but the short bursts extend to lower intensities than seen for long bursts. The average X-ray flux density at 3 keV at 11 hr for the short bursts is $ = 9.6\\times10^{-3}\\mu$Jy, which is more than an order of magnitude less than the average for long bursts of $ = 0.10\\mu$Jy. The X-ray to gamma-ray correlation in Figure 2 has a positive correlation with a slope of roughly unity. This suggests that brighter bursts have more kinetic energy in the afterglow phase to power the afterglow. This is a manifestation of similar radiative efficiency among different bursts and between long and short GRBs. Such a point was made by Zhang et al. (2007) based on an analysis of a smaller sample of early {\\it Swift} GRBs. Except for the bursts below the ``dark'' line, most bursts in Figure 1 are confined between lines with $\\beta_{OX} =$ 0.5 and 1.0. This is consistent with a general interpretation that the optical and X-ray emission belong to the same spectral component with an index close to 0.75. Within the standard model for emission via synchrotron radiation, for slow cooling which is generally relevant at $t = 11$ hr, one expects $\\beta_{OX} \\sim (p-1)/2$ for $\\nu_m < \\nu_O < \\nu_X < \\nu_c$, which has a typical value of 0.75 for electron distribution power law $p = 2.5$. (An equivalent statement is that for this spectrum, the predicted ratio $F_R/F_X \\approx 350$ yields a line intermediate between % $\\beta_{OX}=0.5$ and $1.0$ in % Fig. 1.) % This suggests that on average, the cooling frequency is above or not much below the X-ray band at 11 hr. \\subsection{Dark GRBs} Another comparison of short and long GRBs relates to dark bursts. Jakobsson et al. (2004, see also De Pasquale et al. 2003) used the simple criterion to define dark bursts as those with extremely low optical to X-ray afterglow ratio, falling below the line of optical to X-ray spectral index, $\\beta_{\\rm OX}$, equal to 0.5. It may seem counterintuitive that there can be dark bursts with optical detections and bursts not detected in the optical that are not ``dark'', but the important criterion is how optically faint the burst is relative to its X-ray flux. For the pre-{\\it Swift} sample there were 5 bursts with upper limits below the dark-burst line (restricting the Jakobsson et al. sample to include only those with upper limits fainter than $m_R=23$, or $\\sim2\\mu$Jy), compared to 24 bursts with actual measurements (not upper limits) above the line, giving a fraction of $\\sim17$\\% in the dark category. For {\\it Swift} there are 2 bursts with upper limits (GRB 050713B and 061222A) and 3 cases with measurements (GRB 060210, 070419B and 070508) below the line compared with 34 long bursts above the line for a fraction of $\\sim17$\\% in the dark category, the same as the pre-{\\it Swift} sample. The conclusion is that {\\it Swift} is sampling the same source environments as previous instruments. The discovery of 3 cases of dark bursts with optical detections is particularly interesting. One possible concern with this finding is that {\\it Swift} X-ray afterglows are contaminated in many bursts by emission components not from the external shocks, e.g. X-ray flares. In such instances, the Jakobsson et al. (2004) approach to define dark bursts is no longer relevant since it assumes that the X-ray and optical emission is from the same emission component, but separated by a cooling break. However, the X-ray lightcurves for the {\\it Swift} dark bursts are smooth around 11 hr (and beyond the end of the X-ray plateau), with no significant contamination from other components. These are real ``dark'' bursts from both an observational and physics perspective. Correlation analyses between optical and X-ray can help answer the question of whether these two afterglow components originate from the same physical processes. It is assumed in the Jakobsson et al (2004) study that both X-ray and optical emission arise from the external forward shock. Multiwavelength observations in the {\\it Swift} era reveal puzzling chromatic features of afterglow breaks (e.g., Panaitescu et al. 2006; Liang et al. 2007, 2008) that are not consistent with the simplest forward shock model. Models invoking non-forward-shock origin of X-ray afterglows have been discussed in the literature (e.g., Genet et al. 2007; Uhm \\& Beloborodov 2007; Ghisellini et al. 2007; Shao \\& Dai 2007; Panaitescu 2008). On the other hand, analyses suggest that the X-ray data are generally consistent with the temporal index and spectral index relations (e.g., Zhang \\& M\\'esz\\'aros 2004) predicted by the forward shock models, although not in every case. (Liang et al. 2007, Willingale, et al. 2007). The optical/X-ray data of some bursts (e.g., Grupe et al. 2007; Mangano et al. 2007) are consistent with the same forward shock model. Regardless of the exact process, the analysis presented in this paper shows that generally optical/X-ray afterglow fluxes are correlated, which suggests that they are due to the same emission process. The few cases well below the correlation line are found to be dark due to extinction in the host galaxy. For the first time we can search for dark short bursts. No short bursts are seen that fall below the dark-burst line. It is hard to find dark GRBs using this criterion since X-ray afterglow fluxes are also low for the short bursts. However, there are some short bursts with bright X-ray afterglow, and, to date, none of those is seen to be highly deficient in optical afterglow. Statistics are still small with only 5 optical detections, but if the observed trend continues we will be able to conclude that short bursts do not occur in regions with extremely high extinction as occurs for some long bursts. We are beginning to have optical detections of bursts below the dark burst line. In one of the three dark bursts with detections (GRB 060210), the burst is found to have high extinction associated with its host galaxy, explaining the low optical flux (Curran et al. 2007b). By modeling the differences between $\\beta_{\\rm opt}$, $\\beta_X$, and $\\beta_{OX}$, and taking into account the Lyman$-\\alpha$ absorption ($z=3.91$), the authors find the $R-$band source extinction could be $3.9\\pm0.7$ mag ($\\nu_c > \\nu_O$) or $6.7\\pm0.6$ mag ($\\nu_c < \\nu_O$). This is an important development in our understanding of dark GRBs. (For two of the three dark bursts with detections - 070419B and 070508 - there has not yet been sufficiently detailed follow-up work on the putative hosts for constraints to be placed on the host extinction.) Assuming that the dark bursts can be largely explained by extinction, then the optical - X-ray correlations, ignoring the dark bursts, would hold true. We note that new studies are being done to examine dark burst definitions (van der Horst et al. 2008). \\subsection{Prompt Fluence and Flux Comparisons} The comparison of fluences and peak fluxes in the prompt emission as shown in Figure 3 is a different kind of study than in the other two above. In this case, the strong observed correlation and high degree of separation of short and long bursts is expected; brighter bursts with higher peak fluxes naturally have higher fluences and short bursts tend to have lower fluence for a given flux by the very fact of their short duration. Within the short and long classes, the spread in fluence that is seen for a given peak flux is due to the diversity of durations and spectral indices. Bursts with longer duration and hard spectra have higher fluences for the same peak flux. It is interesting to note in Figure 3 that the short bursts tend to be fluence limited in the BAT, while long bursts tend to be peak flux limited. This is due to the way BAT operates. A valid GRB trigger requires a statistically significant excess in both the rate and image domains (Fenimore et al. 2004). The ability to form an image depends on the number of photons collected on various trigger timescales, which is related to the burst fluence. Even for relatively high peak-fluxes, short bursts can have low fluence values and be limited in the number of photons available for the image trigger. On the other hand, long bursts tend to have higher fluences for a given peak-flux and become rate limited before the image limit is reached. BAT also has a pure-image mode for triggering where very long duration GRBs and other transients are found by comparing sky images instead of having a rate trigger. The lowest long-burst point in Figure 3 at a peak flux of $\\sim0.1$ was such an image-mode trigger for the very long ($T90 = 35$ min) and weak GRB 060218. A caveat on the above discussion is that the BAT trigger algorithm is complex with $\\sim500$ different trigger criteria evaluated. There are many different thresholds and limits coming into play for short and long burst triggering, with some mix of flux and fluence limits for both types. This study was based on a 1 s binning for the gamma-ray fluxes. We have also investigated the effect of using a smaller bin size of 64 ms. Smaller bins pick out larger peak flux values when there is short time structure or when the burst has a duration shorter than the bin size. The effect of the smaller bin size is to shift the short bursts to the right (higher peak flux) relative to the long bursts by about a factor of 5. The larger bin size that we use allows for better statistics and is more reliable for long bursts. In either case, the short bursts tend to cluster toward lower fluences than long bursts." }, "0808/0808.0671_arXiv.txt": { "abstract": "It has recently been pointed out that if the surface tension of quark matter is low enough, the surface of a strange star will be a crust consisting of a crystal of charged strangelets in a neutralizing background of electrons. This affects the behavior of the surface, and must be taken into account in efforts to observationally rule out strange stars. We calculate the thickness of this ``mixed phase'' crust, taking into account the effects of surface tension and Debye screening of electric charge. Our calculation uses a generic parametrization of the equation of state of quark matter. For a reasonable range of quark matter equations of state, and surface tension of order a few MeV/fm${}^2$, we find that the preferred crystal structure always involves spherical strangelets, not rods or slabs of quark matter. We find that for a star of radius 10 km and mass $1.5 M_\\odot$, the strangelet-crystal crust can be from zero to hundreds of meters thick, the thickness being greater when the strange quark is heavier, and the surface tension is smaller. For smaller quark stars the crust will be even thicker. ", "introduction": "\\label{sec:intro} Quarks in their most familiar form are confined in protons and neutrons that make up standard nuclear matter. However, according to the ``strange matter hypothesis'' \\cite{Bodmer:1971we,Witten:1984rs} this form of matter might be metastable and the fully stable state would then be ``strange matter'', which contains roughly equal numbers of up, down, and strange quarks. Large (kilometer-sized) pieces of strange matter are ``strange stars'' (for a review see Ref.~\\cite{Weber:2004kj}); small nuggets of strange matter are ``strangelets'' \\cite{Farhi:1984qu}. The strange matter hypothesis remains a fascinating but unproven conjecture. In this paper we will assume that it is correct, and investigate the structure of the crust of a strange star. The traditional picture of the surface of a strange star is a sharp interface, of thickness $\\sim$ 1 Fermi. Below the interface lies quark matter, the top layer of which is positively charged. Above the interface is a cloud of electrons, sustained by an electric field which could also support a thin nuclear matter crust in suspension above the quark matter~\\cite{Alcock:1986hz,Stejner:2005mw}, as long as the strange star is not too hot~\\cite{Usov:1997eg}. However, if the surface tension $\\si$ of the interface between quark matter and the vacuum is small enough, the surface will take on a much more complicated structure. If $\\si$ is less than a critical value $\\sicrit$ then large lumps of strange matter are unstable against fission into smaller pieces \\cite{Jaikumar:2005ne,Alford:2006bx}. As a result, the simple surface described in the previous paragraph is unstable, and is replaced by a mixed phase involving nuggets of positively-charged strange matter in a neutralizing background of electrons. It is reasonable to guess that the ground state is a regular lattice, leading to a crust with a crystalline structure. (Note that this crystal is completely different from the Larkin-Ovchinnikov-Fulde-Ferrell phase of quark matter \\cite{Alford:2000ze,Casalbuoni:2003wh}, where the quark matter density is uniform, but the pairing gap varies in space.) Jaikumar, Reddy, and Steiner \\cite{Jaikumar:2005ne} conjecture that the strangelet crystal crust will actually be a multi-layer structure of mixed phases, analogous to the ``nuclear pasta'' phases that occur in models of the inner crust of a conventional neutron star \\cite{Ravenhall:1983uh}. At the outer edge of the crust, we expect a dilute low-pressure lattice of small strangelets in a degenerate gas of electrons. As we descend in to the star, the pressure rises, and the structure is modified (becoming denser, and perhaps changing to rods, slabs, cavities, etc). At a critical pressure $\\pcrit$ the mixed phase is no longer stable, and there is a transition to uniform neutral quark matter. As one burrows deeper into the star the pressure continues to rise, and there may be other phase transitions between different phases of quark matter \\cite{Alford:2007xm}, but those will not concern us here. Note that in this scenario the strange star depends on gravity for its existence. In the absence of gravity, it would undergo fission into strangelets. Ref.~\\cite{Jaikumar:2005ne}, assuming zero surface tension and neglecting Debye screening, estimated that the mixed phase crust might be $40-100$~m thick, with $\\pcrit\\approx 1000~\\MeV^4$. This is an interesting result because if a strange star has a sufficiently thick crystalline crust, it might be hard to distinguish from the crust of a neutron star. Astrophysical properties that are sensitive to the crust include cooling behavior, neutrino and photon opacity during a supernova, the photon emission spectrum, glitches, and frequencies of seismic vibrations which are observed after giant flares in magnetars. For further discussion and references see Sec.~\\ref{sec:discussion}. This paper will make a more careful calculation of the properties of a strangelet crystal crust, including the effects of Debye screening \\cite{Heiselberg:1993dc} and surface tension. We expect that the properties of the crust will emerge from a competition between various different contributions to the energy. Charge separation is often favored by the internal energy of the phases involved, because a neutral phase is always a maximum of the free energy with respect to the electrostatic potential (see \\cite{Ravenhall:1983uh,Glendenning:1992vb}; for a pedagogical discussion see \\cite{Alford:2004hz}). The domain structure is determined by surface tension (which favors large domains) and electric field energy (which favors small domains). Debye screening is important because it redistributes the electric charge, concentrating it in the outer part of the quark matter domains and the inner part of the surrounding vacuum, and thereby modifying the internal energy and electrostatic energy contributions. To make an estimate of the thickness of the crust we need to calculate the equation of state of the mixed phase, i.e.~the energy density $\\ep_{\\rm mp}$ as a function of the pressure $p_{\\rm mp}$. The thickness of the crust for a star of mass $M$ is then \\beq \\Delta R = \\Rstar\\Bigl(\\frac{\\Rstar}{GM}-2\\Bigr)\\int_0^{\\pcrit} \\frac{1}{\\ep_{\\rm mp}} dp_{\\rm mp} \\ , \\label{dr} \\eeq in $\\hbar=c=1$ units. This expression follows from the Tolman Oppenheimer Volkoff equation \\cite{Tolman:1939jz,Oppenheimer:1939ne}, assuming that $\\De R\\ll \\Rstar$, and that everywhere in the crust the pressure is much smaller than both the local energy density and the average energy density of the whole star. These are very good approximations for the cases that we study. We obtain $\\ep_{\\rm mp}$ as a function of $p_{\\rm mp}$ by dividing the strangelet lattice into unit cells (``Wigner-Seitz cells'') and calculating the pressure at the edge of a cell as a function of its energy density. We study cells that are three-dimensional (a lattice of strangelets in a degenerate gas of electrons), two-dimensional (rods of strange matter in a degenerate gas of electrons) and one-dimensional (slabs of strange matter interleaved with regions of degenerate electron gas). Our approach is similar to that used in studying mixed phases of quark matter and nuclear matter in the interior of neutron stars \\cite{Maruyama:2007ey}. We build on the formalism for a generic quark matter equation of state and infinitely-large Wigner-Seitz cells that was developed in \\cite{Jaikumar:2005ne,Alford:2006bx}. The main assumptions that we make are:\\\\ 1) Within each Wigner-Seitz cell we use a Thomas-Fermi approach, solving the Poisson equation to obtain the charge distribution, energy density, and pressure. This is incorrect for very small strangelets, where the energy level structure of the quarks becomes important \\cite{Madsen:1994vp,Amore:2001uf}; such corrections may be relevant for the very low pressure (outer crust) part of our results (see Sec.~\\ref{sec:discussion}). \\\\ 2) We assume our $D$-dimensional Wigner-Seitz cells to be $D$-spheres. In reality the cells will be unit cells of some regular lattice (cubic, hexagonal close packed, etc). However, as long as the cell is much bigger than the strangelet inside it, we expect this approximation to be reasonably accurate. We will only report results for cases where $\\Rcell>2R$ ($R$ being the radius of the quark matter in the center of the cell, which we expect will have a rotationally symmetric shape because of the surface tension). In some cases this assumption is violated, and we will then only be able to obtain a lower limit on the crust thickness (see Sec.~\\ref{sec:thickness-results}). \\\\ 3) We treat the interface between quark matter and the vacuum as a sharp interface, with no charge localized on it, which is characterized by a surface tension. We neglect any surface charge that might arise from the reduction of the density of states of strange quarks at the surface \\cite{Madsen:2000kb,Madsen:2001fu,Madsen:2008bx,Oertel:2008wr}. We also neglect the curvature energy of a quark matter surface \\cite{Christiansen:1997rc,Christiansen:1997vt}, so we do not allow for ``Swiss-cheese'' mixed phases, in which the outer part of the Wigner-Seitz cell is filled with quark matter, with a cylindrical or spherical cavity in the center, for which the curvature energy is crucial. Note that these phases would be expected to occur at higher pressure than the ones we study, so including them is likely to make the crust even thicker than we predict. \\\\ 4) We assume that the chemical potential for negative electric charge $\\mue$ is much less than the chemical potential for quark number $\\mu$. This allows us to expand the quark matter equation of state in powers of $\\mue$, and means that within the quark matter we can ignore the contribution of electrons to the charge and pressure. This is a very good approximation for small strange quark mass, which corresponds to small $\\nQ$ in our parameterization. For the largest value of $\\nQ$ that we study, $\\mue$ in neutral quark matter is close to 100 MeV, and the assumption is still reasonable.\\\\ 5) We assume that only electrons are present, with no muons. This is valid as long as $\\mue$ is less than the muon mass $m_\\mu$, which is true for all the cases that we study.\\\\ 6) We assume that $\\mue$ is always much greater than the electron mass. Thus in the degenerate electron gas, we can take the electrons to be massless, which simplifies the Thomas-Fermi calculation of their charge distribution. Since $\\mue$ drops monotonically from the center of the cell to its edge, this condition will only be violated for very large cells (very low pressures).\\\\ 8) We always work at zero temperature. The temperature of the surface of a compact star, even during a flare \\cite{Lyubarsky:2002cs}, is expected to be less than 100 keV, so we expect this to be a reasonable approximation. ", "conclusions": "\\label{sec:discussion} The calculations described in this paper give us a more precise picture of the strangelet-crystal crust of a quark star. The results presented in Table~\\ref{tab:crusts} show that the thickness of the strangelet-crystal crust of a strange star is very sensitive to the surface tension $\\si$ of the interface between quark matter and the vacuum, and to the quark matter parameters $\\nQ$ and $\\chiQ$ \\eqn{eqn:generic_EoS}, which determine the response of the quark matter to deviations of the electrostatic potential from its neutrality value. Our results are compatible with those of Ref.~\\cite{Jaikumar:2005ne} where an upper limit on the crust thickness was obtained by ignoring surface tension and Debye screening. The crust is thickest for large $\\nQ$ and small $\\chiQ$ (large $\\la_D$). As discussed in Sec.~\\ref{sec:thickness-results}, we find that values of surface tension in the physically expected range, around 1 to $10~\\MeV\\fm^{-2}$, reduce the thickness of the crust and may even eliminate it completely, but it remains possible that a quark star of radius 10~km could have a crust several hundred meters thick. From Table~\\ref{tab:factors} we see that for a smaller star the crust could be even thicker. The geometry of the mixed phase in our crusts, on the other hand, shows no variation at all. It is always three-dimensional, containing spherical droplets of quark matter. We never find any case where a two-dimensional (rod) or one-dimensional (slab) geometry is favored. Our calculations and results suggest two directions for future work: firstly, one could study phenomenological consequences of our understanding of the strangelet-crystal crust of a quark star. Secondly, one could improve on our treatment of the strangelet crystal, by relaxing some of the assumptions listed in Sec.~\\ref{sec:intro}. The most obvious phenomenological task is to revisit computations of the frequencies of seismic vibrations which are observed after giant flares in magnetars \\cite{Watts:2006hk,Chugunov:2006kk}. Ref.~\\cite{Watts:2006hk} found that the strangelet crystal crust did not have the right spectrum of toroidal shear modes to account for current observations: it would be interesting to see whether taking in to account the surface tension and Debye screening affects that conclusion. Other aspects of the phenomenology of the crust could also be studied, for example (a) the thermal response of the crust to accretion \\cite{2000ApJ...531..988B}; (b) the role of the crust in the trapping of neutrinos and photons just after a type-II supernova \\cite{Burrows:2004vq}; (c) the spectrum of photons radiated from the surface of a quark star \\cite{Page:2002bj,Jaikumar:2004zy,Harko:2004ts}; (d) the contribution of the crust to the moment of inertia and glitches \\cite{1992ApJ...400..647G}; (e) the damping of $r$-modes in by shear viscosity in the crust \\cite{2000ApJ...529L..33B,Caballero:2008tx} (for quark stars, the contribution from the interior has been calculated \\cite{Madsen:1999ci,Jaikumar:2008kh}); (f) the thermal relaxation time of the crust and its response to the post-supernova ``cooling wave'' \\cite{Gnedin:2000me}. The thermal relaxation time of the crust depends on the thermal conductivity, for which we can make a very rough estimate using appendix A of Ref.~\\cite{Gnedin:2000me}. We find values of order a few hundred $\\MeV^2$ at $T\\sim 0.1~\\MeV$, which is comparable to the range $10^{18}~{\\rm erg}\\,{\\rm cm}^{-1}{\\rm s}^{-1}{\\rm K}^{-1}$ for low-density nuclear matter (Ref.~\\cite{Gnedin:2000me}, Fig.~4). We defer a more accurate calculation to future work. To improve on our treatment, the most pressing issues are to use a realistic shape for the Wigner-Seitz cells (which should be unit cells of some regular lattice, rather than spheres), to include shell-model corrections for the smallest strangelets, and to allow for ``Swiss-cheese'' phases where most of the unit cell consists of quark matter, with a hole at the center containing electrons. As discussed in Sec.~\\ref{sec:thickness-results}, the shape of the cell becomes important at very high pressures, and our use of the spherical approximation meant that in some cases we could only obtain lower limits on the crust thickness. Studying more realistic shapes is straightforward in principle, but would require a more demanding multidimensional numerical solution of the Poisson equation. Shell-model corrections can be of order one MeV per quark for strangelets of size $R\\lesssim 5~\\fm$ \\cite{Madsen:1994vp,Amore:2001uf}, which is not negligible relative to our typical enthalpy per quark (Fig.~\\ref{fig:stability-search}), and may therefore affect our results for the outer part of the crust, where we predict strangelets as small as 3~fm (Fig.~\\ref{fig:crust}). Treating Swiss-cheese phases would require us to include curvature energy as well as surface tension. This highlights the fact that we treated the interface between quark matter and vacuum in a very simplified way, as a zero-width interface with a surface tension. However, since the quark confinement distance is about 1~fm, the interface might well have structure on this distance scale. Like the shell-model effects described above, this could be relevant to the low-pressure regime, where the strangelets can be as small as a few fm. There are even indications that when such physics is taken into account, the CFL phase may undergo some degree of charge separation \\cite{Madsen:2001fu,Oertel:2008wr}, raising the possibility that there might be some sort of crystalline crust on quark stars made of CFL quark matter." }, "0808/0808.2168_arXiv.txt": { "abstract": "We report the discovery of six infrared stellar-wind bowshocks in the Galactic massive star formation regions M17 and RCW49 from {\\it Spitzer} GLIMPSE (Galactic Legacy Infrared Mid-Plane Survey Extraordinaire) images. The InfraRed Array Camera (IRAC) on the {\\it Spitzer Space Telescope} clearly resolves the arc-shaped emission produced by the bowshocks. We combine {\\it Two Micron All-Sky Survey (2MASS),} {\\it Spitzer,} {\\it MSX,} and {\\it IRAS} observations to obtain the spectral energy distributions (SEDs) of the bowshocks and their individual driving stars. We use the stellar SEDs to estimate the spectral types of the three newly-identified O stars in RCW49 and one previously undiscovered O star in M17. One of the bowshocks in RCW49 reveals the presence of a large-scale flow of gas escaping the \\hii\\ region at a few $10^2$ \\kms. Radiation-transfer modeling of the steep rise in the SED of this bowshock toward longer mid-infrared wavelengths indicates that the emission is coming principally from dust heated by the star driving the shock. The other 5 bowshocks occur where the stellar winds of O stars sweep up dust in the expanding \\hii\\ regions. ", "introduction": "The Solar wind ends in a termination shock \\citep[e.g.][]{V105}, where the pressure of the heliosphere balances the ram pressure of the surrounding interstellar medium (ISM). Massive stars with more energetic winds generate much stronger shocks. In cases where the relative motion between the star driving the wind and the ambient ISM is large, the shock will be bent back around the star. If the relative velocity is supersonic, the ambient ISM gas is swept into a second shock, forming an arc-shaped ``bowshock'' that is separated from the termination shock by a contact discontinuity. Stellar-wind bowshocks have been reported for a variety of sources, including nearby runaway O stars \\citep{vBu88,vBu95,Nor97,BB05,CP07,FML07}, high-mass X-ray binaries \\citep{EC92,K97,HK02}, LL Ori-type stars \\citep{Bally00}, radio pulsars \\citep{GS06}, Galactic center O stars \\citep{geb04,geb06}, and mass-losing red giants \\citep{Martin07}. Recently, an infrared (IR) bowshock has been observed around the young A-type star $\\delta$ Vel \\citep{AG08}. Cometary \\hii\\ regions also resemble bowshocks, due either to density gradients in the ambient gas or to motion of the ionizing source with respect to the interstellar surroundings \\citep{vB90,AH06}. Both the direction of a bowshock and its ``standoff distance'' from the star generating the wind are determined by the velocity of the star with respect to the surrounding medium. In the case of runaway O stars, this is dominated by the high space motion of the star. We report the detection of three mid-IR bowshocks in each of two massive star formation regions: M17 and RCW49. Two of the bowshocks in M17 are around known O stars. We will demonstrate that the other bowshocks are also around likely O stars. Since these stars are in or near expanding \\hii\\ regions, we find that the direction of the bowshock is determined principally by the flow of the ISM rather than the space motion of the star. ", "conclusions": "We have observed 6 prominent IR bowshocks in M17 and RCW49. These objects appear to be produced by the winds of individual O stars colliding with large-scale interstellar gas flows in the \\hii\\ regions. One bowshock, M17-S3, may be the leading edge of an evaporating globule containing a newly-formed and previously undiscovered O star in the well-studied M17 region. All three bowshocks associated with RCW49 lead us to identify new candidate O stars. Our stellar classifications also suggest that the true distance to RCW49 is less than the kinematic distance of 6 kpc. The bowshocks are bright at IR wavelengths due to emission from dust swept up from the ambient ISM and heated by radiation from the bowshock driving stars. As \\citet{AG08} note, IR excess emission from a bowshock could be attributed to the presence of a circumstellar disk, particularly when the bowshock morphology is not spatially well-resolved. This can be a pitfall for observational studies of accreting massive protostars. The collective winds of the most luminous stars in young, massive clusters produce overlapping large-scale flows that hollow out thermally hot cavities in the parent molecular cloud \\citep{T03}. The largest bowshock presented here, RCW49-S1, is evidence that the combined winds of the ionizing stars in Westerlund 2 have escaped the \\hii\\ region, creating a flow of hot gas moving at a few $10^2$ \\kms\\ that extends at least 16 pc away from RCW49. The driving stars of the other 5 bowshocks are surrounded by ionized gas and dust of their natal \\hii\\ regions, where the density of the ambient medium ($n_0\\sim 10^3$ cm$^{-3}$) is sufficiently high to produce the observed bowshocks with a relative velocity of only 10--20 \\kms. The winds of the bowshock driving stars do not directly encounter the ${>}2000$ \\kms\\ winds from the most massive stars in the cluster. Instead, the bowshocks are shaped by the expansion of the ionized gas in the \\hii\\ regions relative to the orbital motions of the stars. Eventually, supernova explosions will produce high velocity shock waves that heat and disperse the original gas cloud. In star forming regions like M17 and RCW49 that have not yet been disrupted by supernovae, IR bowshocks serve as interstellar ``weather vanes,'' indicating the speed and direction of large-scale gas flows at points within and around giant \\hii\\ regions." }, "0808/0808.1111_arXiv.txt": { "abstract": "{The exploitation of clusters of galaxies as cosmological probes relies on accurate measurements of their total gravitating mass. X-ray observations provide a powerful means of probing the total mass distribution in galaxy clusters, but might be affected by observational biases and rely on simplistic assumptions originating from our limited understanding of the intracluster medium physics.} {This paper is aimed at elucidating the reliability of X-ray total mass estimates in clusters of galaxies by properly disentangling various biases of both observational and physical origin.} {We use N-body/SPH simulation of a large sample of $\\sim 100$ galaxy clusters and investigate total mass biases by comparing the mass reconstructed adopting an observational-like approach with the true mass in the simulations. X-ray surface brightness and temperature profiles extracted from the simulations are fitted with different models and adopting different radial fitting ranges in order to investigate modeling and extrapolation biases. Different theoretical definitions of gas temperature are used to investigate the effect of spectroscopic temperatures and a power ratio analysis of the surface brightness maps allows us to assess the dependence of the mass bias on cluster dynamical state. Moreover, we perform a study on the reliability of hydrostatic and hydrodynamical equilibrium mass estimates using the full three-dimensional information in the simulation.} {A model with a low degree of sophistication such as the polytropic $\\beta$-model can introduce, in comparison with a more adequate model, an additional mass underestimate of the order of $\\sim 10 \\%$ at $r_{\\mathrm{500}}$ and $\\sim 15 \\%$ at $r_{\\mathrm{200}}$. Underestimates due to extrapolation alone are at most of the order of $\\sim 10 \\%$ on average, but can be as large as $\\sim 50 \\%$ for individual objects. Masses are on average biased lower for disturbed clusters than for relaxed ones and the scatter of the bias rapidly increases with increasingly disturbed dynamical state. The bias originating from spectroscopic temperatures alone is of the order of $10 \\%$ at all radii for the whole numerical sample, but strongly depends on both dynamical state and cluster mass. From the full three dimensional information in the simulations we find that the hydrostatic equilibrium assumption yields masses underestimated by $\\sim 10-15 \\%$ and that masses computed by means of the hydrodynamical estimator are unbiased. Finally, we show that there is excellent agreement between our findings, results from similar analyses based on both Eulerian and Lagrangian simulations, and recent observational work based on the comparison between X-ray and gravitational lensing mass estimates.} {} ", "introduction": "\\label{intro.sec} Galaxy clusters are the largest virialized structure known in the universe. According to the hierarchical clustering model of structure formation they form by the gravitational collapse of the rare peaks of the primeval density field, on scales of the order of $\\sim 10$ Mpc. Within this scenario their formation and evolution is a sensitive function of the cosmological matter density parameter $\\Omega_{\\mathrm{m}}$ and the mass fluctuation amplitude $\\sigma_\\mathrm{8}$, where $\\sigma_\\mathrm{8}$ is the rms linear fluctuation on scales $8 \\, h^{\\mathrm{-1}}$ Mpc. Measurements of their evolution rate can be used to asses the growth in mass of such structures, thereby providing a powerful method to constrain the geometry and matter content of the universe \\citep[see][and references therein]{vo05}. Moreover, because of the spatial extent of the collapse scale, the cluster baryonic fraction $f_{\\mathrm{b}}$ is expected to be close to the cosmic value $\\Omega_{\\mathrm{b}}/ \\Omega_{\\mathrm{m}}$. Measurements of $f_{\\mathrm{b}}$ at high redshifts can be used to derive constraints on the equation of state of the dark energy \\citep{hai01,maj04,al08}. The importance of cluster of galaxies as cosmological probes will be further strengthened with the upcoming high redshift surveys. This is of particular relevance in the new era of precision cosmology, in which studies of cluster evolution will provide independent tests with which to compare constraints on cosmological models extracted from observations of the cosmic background radiation \\citep[e.g.][]{spe07} and distance measurements of high redshift supernovae \\citep[e.g.][]{ton03,rie04,rie07}. From the scenario here outlined it follows that in order to use cluster of galaxies as cosmological probes it is crucial to accurately measure their baryonic and total gravitating mass. The methods used to derive cluster masses are mainly based on the velocity dispersion of the optical galaxy populations \\citep[e.g.][]{biv03,rin03}, observations of the X-ray emitting intracluster medium (ICM) \\citep[e.g.][]{fin01,rei02,et02,arn05,vikhlinin2006} and on gravitational lensing \\citep[e.g.][]{smi02,mahdavi08}. Accurate mass estimates derived from X-ray data are based on the assumptions that both the total potential and the ICM distribution are spherically symmetric and that the ICM is in hydrostatic equilibrium in the cluster potential well. The latter assumption is usually justified by the fact that the estimated ICM sound crossing times are short when compared against cluster ages \\citep{sar86}. Under these assumptions the ICM is a faithful tracer of the underlying matter distribution and the total mass profile can be determined from the gas radial density and temperature profiles. The density profile is recovered from deprojecting of the surface brightness profile, as measured from X-ray flux maps, whereas knowledge of the temperature profile requires the availability of spatially resolved spectroscopy. High quality data taken by the present generation of X-ray telescopes, {\\it Chandra} and {\\it XMM-Newton}, allows for nearby clusters accurate measurements of these quantities out to a large fraction of the cluster virial radius \\citep{ma98,vi05,pi05,san06,pra07,leccardi08}. The reliability of cluster mass estimated through the X-ray method can be accurately studied using N-body/hydrodynamical simulations. The principal benefit over analytical methods is that the gas evolution can be treated self-consistently. The validity of the numerical approach is supported by X-ray observations, which show the existence of complex thermal structures and of merging activity \\citep[see][for a review]{ma07}. Since early pioneering studies, hydrodynamical simulations have become a widely used tool to investigate cluster formation and evolution in different cosmological scenarios \\citep[cf.][]{vo05}. The numerical resolution of the simulations and the modelization of the cluster gas physics has been improved over the years. The latter now incorporates radiative cooling \\citep{yos00,lew20,per20,mua02,dav02}, metal enrichment of the ICM by supernovae and energy feedback \\citep{kay03,tor03,valda03,bor04,kay04,kra05}. The accuracy of cluster X-ray mass estimators has been tested by means of N-body/hydrodynamical simulations in a variety of papers \\citep{evr90,evr96,kay04,rasia06,kay07,nagai07}. \\cite{evr90} first pointed out the existence of a significant bias in the binding mass estimates when using the isothermal $\\beta-$model. \\cite{evr96} confirmed that the source of this discrepancy is related to the isothermal and hydrostatic assumptions. The lack of validity of the hydrostatic assumption is observationally motivated by optical and X-ray maps, which show the existence of substructure with an ongoing merger activity, and is numerically supported by a number of authors \\citep[e.g.][]{kay04,ras04,nagai07}, who found that in the simulations the ICM is not perfectly in hydrostatic equilibrium. This implies the presence of residual gas bulk (laminar) and turbulent (random kinetic) motions, and leads to an underestimate of the masses because of additional non-thermal pressure support which is not accounted for by the hydrostatic equilibrium equation. With respect the isothermal $\\beta-$model the modelization of the ICM has been significantly improved \\citep[e.g.][]{vikhlinin2006} with observational progresses, which showed the existence of temperature profiles declining with radius \\citep{de02,vi05,pi05,san06,pra07}. These features are well reproduced out of the core radii in hydrodynamical simulations which incorporate cooling and feedback \\citep{mua02,kay03,tor03,valda03,bor04}. In order to properly assess the reliability of cluster X-ray mass estimators it is however necessary to construct mock observations of simulated clusters which must reproduce spectroscopic measurements as expected from X-ray telescopes. This is motivated by the presence of complex thermal structures in the ICM, which bias the (measured) spectral fit temperatures towards lower values than the average emission weighted cluster temperatures defined theoretically \\citep{ma01,mazzotta04,valda06}. The dependence of X-ray mass estimators on spectral biases and other systematics has been investigated through hydrodynamical simulations by a number of authors \\citep{rasia06,kay07,nagai07,je07}. \\cite{nagai07} argued that mass estimates are biased low ($\\sim 5-20 \\%$) even for clusters identified as relaxed. In order to properly disentangle observational biases which arise from spectroscopic measurements from those due to incomplete relaxation of the gas or from the ones caused by an inaccurate modeling of the radial profiles, it is however necessary to derive X-ray mass estimates from a large sample of simulated clusters. This is the main goal of this paper, in which we apply different X-ray mass estimators to a large set of high-resolution hydrodynamical simulations of galaxy clusters. The physics of the gas includes radiative cooling, star formation, chemical enrichment and energy feedback. The sample comprises $\\sim 100$ clusters, the size of the sample being a critical quantity in order to extract sub-samples large enough to give a meaningful statistics. We discuss the dependence of the mass bias at different radii upon the adopted analytical models and the chosen radial range used to perform the fits of the profiles, the cluster dynamical state as well as the impact on the mass bias which follows from the use of spectral temperatures. As a statistical indicator to quantify the cluster dynamical state through the analysis of X-ray maps we adopt the power ratio method \\citep[see][and references therein]{je05}. We also investigate the limit of applicability of the dynamical equilibrium equation when used to recover the cluster true masses in the presence of significant non-thermal gas pressure. The paper is organized as follows. In Sect. \\ref{simulations.sec} we present the procedure for simulating the cluster sample. In Sect. \\ref{observables.sec} we describe how we generate and analyze the synthetic X-ray observations that are used in Sect. \\ref{mass.sec} to recover the total mass distribution. In Sect. \\ref{masssim.sec} we investigate the reliability of the hydrostatic and hydrodynamical mass estimators from the full three-dimensional information provided by the simulations. Finally, we discuss our main results and present our conclusions in Sect. \\ref{conclusions.sec} . ", "conclusions": "\\label{conclusions.sec} This work is aimed at elucidating the reliability of X-ray total mass estimates in clusters of galaxies using N-body/SPH simulation of a large sample of clusters. The physical modeling the gas includes radiative cooling, star formation, supernovae heating, and metal enrichment. The large number of simulated clusters enables us to derive very robust conclusions through a statistical analysis of the sample. The total mass is recovered adopting an observational-like approach and compared with the true mass in the simulations. Surface brightness and temperature profiles, that we generate from the simulations, are used to estimate the cluster mass at different overdensities ($r_{\\mathrm{2500}}, r_{\\mathrm{500}}$, and $r_{\\mathrm{200}}$) by means of the hydrostatic equilibrium equation. We explore various models and conditions under which the mass is reconstructed in order to entangle different mass biases. In addition, a power ratio analysis of the surface brightness maps allows us to assess the dependence of the mass bias on cluster dynamical state. Moreover, we perform a study on the reliability of hydrostatic and hydrodynamical equilibrium mass estimates using the full three-dimensional gas distribution in the simulation. In the following we list and discuss our main findings. \\begin{enumerate} \\item Our analysis shows that it is very important to use analytical models with a large amount of parametric freedom when modeling the shape of the ICM temperature and surface brightness radial profiles. A model with a low degree of sophistication such as the polytropic $\\beta$-model can introduce a very large {\\it modeling bias}. Compared to the more sophisticated extended $\\beta$-model, which is found to be extremely accurate in following the slope changes of the gas profiles, it additionally leads to average mass underestimates of the order of $\\sim 5$, $10$, and $15\\%$ at $r_{\\mathrm{2500}}$, $r_{\\mathrm{500}}$, and $r_{\\mathrm{200}}$, respectively. \\item The bias originating from extrapolating of the mass profiles beyond the radial range probed by observation can be extremely large. We find that the underestimate from extrapolation alone is of the order of $\\sim 10 \\%$ at $r_{\\mathrm{200}}$, and lower at smaller cluster-centric distances, when considering an average over the whole sample. However, for individual objects, the {\\it extrapolation bias} can be as large as $\\sim 50 \\%$. \\item The unrelaxed {\\it dynamical state} of a cluster can also lead to mass underestimates. The total mass is on average biased lower for disturbed clusters than for the relaxed ones. Furthermore, we find that the bias values are much more scattered around the mean for disturbed objects than for the relaxed ones. If mass-weighted temperature profiles are adopted, the mean mass bias is at most $\\sim -5 \\%$ at all cluster-centric distances for relaxed clusters, but can be as large as $\\sim -20 \\%$ for the most disturbed ones. Our results are in excellent agreement with the findings of \\cite{je07}, as shown by the comparison of Fig. \\ref{fig:fig1} in this work and Fig. 9 (top panel) in their paper, where the correlation between the same quantities are shown. This agreement is of particular importance, considering that the analysis by \\cite{je07} is based on simulations performed with the adaptive mesh refinement code {\\it Enzo}. Our findings are also in good agreement with the results of \\cite{kay07}. \\item Mass estimates derived using spectroscopic temperatures are lower that those derived from mass-weighted temperatures (i.e. the true gas temperatures). We find that the {\\it spectroscopic temperature bias}, i.e. the bias originating from spectroscopic temperatures alone, is of the order of $-10 \\%$ for the whole numerical sample. A similar value is found by \\cite{kay07}. Mass underestimates derived adopting spectroscopic-like temperature profiles ($7$ and $17 \\%$ on average at $\\Delta=2500$ and 500 for the whole sample) are in good agreement with the values found by \\cite{nagai07} ($12$ and $16 \\%$ are the corresponding values), whose results are based on mock X-ray observations derived from simulations performed with an Eulerian code. We also find very good agreement by performing the same comparison for relaxed and unrelaxed clusters separately. We notice that total mass biases derived from the simulations by \\cite{nagai07} and adopting mass-weighted profiles are on average $-7 \\%$ at $r_{\\mathrm{500}}$ for relaxed clusters \\citep{lau07}. The corresponding value that we find from our Lagrangian simulations is $-5 \\%$. Considering the different nature of the numerical codes and the different implementation of the various physical processes, the agreement is extremely relevant. Our results are in tension with the findings of \\cite{rasia06}, who find, on average, much stronger biases. We notice, however, that a fair comparison is not possible since the work by \\cite{rasia06} focussed only on 5 clusters. Our analysis also shows that the spectroscopic bias depends on dynamical state. We find that the spectroscopic bias is rather small ($-2,-5$, and $-7 \\%$ at $\\Delta= 2500, 500$ and 200) for the most relaxed clusters and of the order of $-12 \\%$ for the most disturbed ones. \\item While the mass bias derived from mass-weighted temperature profiles does not depend on true cluster mass, the one derived from spectroscopic-like temperature exhibits a strong {\\it mass dependence}. For clusters with $M(1.8 \\times 10^{14} \\, h^{-1 }M_{\\odot}$ we find that the spectroscopic bias is larger: e.g. $-7 \\%$ and $-12 \\%$ for the relaxed clusters at $r_{\\mathrm{500}}$ and $r_{\\mathrm{200}}$, respectively. \\item Even in the {\\it ideal case}, i.e. when the mass of relaxed clusters is estimated without extrapolation and adopting a very accurate model for the ICM density and temperature profiles, total masses are affected by the spectroscopic temperatures. In this case the mean total mass bias is $\\sim (-4,-3)$, (-12,-9), and $(-15,-3) \\%$ at $r_{\\mathrm{2500}}, r_{\\mathrm{500}}$, and $r_{\\mathrm{200}}$ for the (most, least) massive clusters in our simulated sample. \\item Even if we assume that the true (mass-weighted) ICM temperature can be reconstructed, masses are biased low (by -2, -9, and $-12 \\%$ at $\\Delta=2500$, 500, and 200, respectively, when taking the average over the whole sample). \\item The latter finding prompted us to investigate the possible {\\it violation of the hydrostatic equilibrium} by using the full, three-dimensional information provided by the simulations. In agreement with the results from our observational-like analysis, we find that the hydrostatic equilibrium assumption yields masses underestimated by $\\sim 10-15 \\%$. This implies that the origin of the bias is not originated by the X-ray reconstruction method. The same level of mass bias has been found by \\cite{ras04} and \\cite{burns07}. \\item In order to elucidate the origin of this bias we compute masses using both {\\it hydrostatic and dynamical equilibrium} mass estimators. The dynamical equilibrium takes into account for both thermal and non-thermal pressure of the ICM. We find that the mass estimated through dynamical equilibrium are on average well estimated. The mean bias profiles of the \"feedback\" clusters simulated by \\cite{kay04} are in very good agreement with our average values at all overdensities. \\item We explore the conditions of applicability of both estimators, i.e. spherical symmetry (torque parameter $\\tau_q\\ll1$) and small radial streaming velocities (radial Mach number $|/c_s| \\ll1$). We find that the biases $b_{HE}$ and $b_{DE}$ do not depend on cluster true mass and that their scatter decreases with decreasing $\\tau_q$ and $|/c_s|$. \\item For clusters with small values of $\\tau_q < 0.15$ and $|/c_s| < 0.1$ we find that the average mass underestimate found from the hydrostatic equilibrium estimator ($5 \\%$ at $r_{\\mathrm{2500}}$ and $10 \\%$ at $r_{\\mathrm{500}}$ and $r_{\\mathrm{200}}$) is extremely well corrected by adopting the dynamical equilibrium estimator. \\end{enumerate} Our results of course depend on the physics included in the simulation. Our modeling of the gas physical processes is incomplete and does not include, for example, energy feedback from active galactic nuclei (AGN). The inclusion of more realistic physics is particularly relevant in the cluster cores, since it is supposed to provide additional heating to the gas which could help to solve the cooling flow problem as well as the overcooling of baryons actually present in current simulations \\citep[e.g.][]{vo05}. The effect on the thermal status of the ICM of energy feedback from the AGN are however not expected to modify in a significant way the ICM properties outside the cluster cores. The validity of this assumption is justified from the radial behavior of the measured temperature profiles, which are in good agreement at $ r \\gtrsim 0.1 \\times r_\\mathrm{200}$ with the present simulations \\citep{valda06} and the ones of other authors \\citep[e.g.][]{nagai07b}. Importantly, we find that, in addition to cold gas clumps, also the diffuse cold gas component which is left after the removal of all resolved clumps substantially biases spectroscopic temperatures low. The good agreement of our temperature profiles with those found by other authors using different codes suggests that the amount of the cool gas component present in our simulations is not affected by numerical issues, There is another issue, which is connected to the gas physical modeling in the simulations, that could be potentially relevant for our analysis. In the runs performed here the artificial viscosity is treated according to the standard SPH formulation \\citep{mo05}, which is a numerical scheme comparatively viscous \\citep{momo97} and inadequate to follow the development of fluid turbulence. It has been shown that the level of kinetic energy in random gas motion could be as high as $ \\sim 30 \\%$ of the gas thermal energy \\citep{dolag05,vazza06}, when the numerical viscosity scheme is generalized as in \\cite{momo97} to properly treat fluid turbulence. It is worth noticing that very high levels of random motions are inconsistent with the upper limits recently derived from the comparison of X-ray and weak lensing mass estimates \\citep{mahdavi08,zhang08}. Direct measurements of the various sources of non-thermal pressure (gas motions, cosmic rays, magnetic fields, etc.) are extremely difficult, and the comparison between X-ray and lensing masses presently provides the most significant constraints on the level of non-thermal pressure in the ICM. \\cite{zhang08} obtain a ratio of $1.09 \\pm 0.08$ between the weak lensing and X-ray mass estimates extracted at $r_\\mathrm{500}$ from a sample of 19 clusters. At the same overdensity, \\citep{mahdavi08} find that the ratio between X-ray and lensing is $0.78 \\pm 0.09$ ($0.85 \\pm 0.10$ after correction for excess structure along the line of sight) for a sample of 18 clusters. The fairly good agreement between these recent observational results and the values found in the present analysis and other similar studies therefore implies a low level of turbulence present in the ICM, thereby suggesting that the gas physics outside the cluster cores is well approximated by the simulations presented here. While we have shown that at present both theoretical models and observations seem to indicate a deviation from the hydrostatic equilibrium of the order of $ \\sim 10 \\%$, further investigations are needed in order to draw robust conclusions on this important issue. Future measurements of X-ray and lensing masses for large samples and measurements of the ICM velocity structure will provide valuable information on the validity of the hydrostatic equilibrium and therefore on the reliability of X-ray clusters as cosmological probes. Moreover, these measurements will also be of particular relevance for constraining gas physics in galaxy clusters. The amount of physical viscosity is in particular a key parameter which quenches the level of turbulence present in the ICM. Physical viscosity has been incorporated in hydrodynamic simulations of galaxy clusters by \\cite{si06} and it was shown that even a modest amount of physical viscosity has significant consequences on ICM properties. Therefore X-ray mass estimates, extracted from a statistically meaningful sample of hydrodynamical SPH simulations which include physical viscosity, could be profitably used to indirectly measure the level of ICM viscosity when contrasted against X-ray and weak lensing mass measurements. These constraints will also likely have a significant impact on those scenarios in which the cooling flow problem is solved by providing additional heating to the gas through energy dissipation." }, "0808/0808.3925_arXiv.txt": { "abstract": "{Nanoflares are small impulsive bursts of energy that blend with and possibly make up much of the solar background emission. Determining their frequency and energy input is central to understanding the heating of the solar corona. One method is to extrapolate the energy frequency distribution of larger individually observed flares to lower energies. Only if the power law exponent is greater than 2 is it considered possible that nanoflares contribute significantly to the energy input. } {Time sequences of ultraviolet line radiances observed in the corona of an active region are modelled with the aim of determining the power law exponent of the nanoflare energy distribution.} {A simple nanoflare model based on three key parameters (the flare rate, the flare duration, and the power law exponent of the flare energy frequency distribution) is used to simulate emission line radiances from the ions \\FeXIX, \\CaXIII, and \\SiIII, observed by SUMER in the corona of an active region as it rotates around the east limb of the Sun. Light curve pattern recognition by an Artificial Neural Network (ANN) scheme is used to determine the values. } {The power law exponents, $\\alpha\\approx2.8$, $2.8$, and 2.6 are obtained for \\FeXIX, \\CaXIII, and \\SiIII\\ respectively. } {The light curve simulations imply a power law exponent greater than the critical value of 2 for all ion species. This implies that if the energy of flare-like events is extrapolated to low energies, nanoflares could provide a significant contribution to the heating of active region coronae.} ", "introduction": "Heating the corona by the dissipation of current sheets was first suggested by \\citet{Gold64} and later developed to form the basis of the nanoflare heating model by \\citet{Levine74} and \\citet{Parker83, Parker88}. The idea is that current sheets arise spontaneously in coronal magnetic fields that are braided and twisted by random photospheric footpoint motions. These current sheets dissipate in many small-scale reconnection events, heating and accelerating plasma in the coronal loops. In the corona, they would give rise to multiple unresolvable loop strands with specific observable signatures \\citep{Zirker94, WWH02, CK04, PK05}. Recently \\citet{Aschwanden08} found evidence against such multi-temperature strands in TRACE coronal images. He concludes that nanoflare heating is only possible if it occurs in the chromosphere/transition region where heating across magnetic field lines can produce the isothermal loops seen in the corona. Irrespective of where the nanoflare energy input sites are, a key question is whether the energy of nanoflares is sufficient to heat the corona or not. Most of the individual nanoflares would be too small to detect and the majority would be small fluctuations on the overall background. That background could be produced by the blending of many small events. The approach taken to estimate their contribution has been to extrapolate the energy frequency distribution of detectable flare-like events. The energy frequency distribution of larger flares tends to follow a power law distribution \\begin{equation} {dN\\over dE} \\sim E^{-\\alpha}, \\end{equation} where $dN$ is the number of flares per energy interval $dE$. The energy of small flares dominates if $\\alpha>2$ \\citep{Hudson91}. This is therefore a critical parameter for the nanoflare heating model. The standard method to determine $\\alpha$ is to evaluate the energy of many flares in a series of observations and then plot their frequency in bins of energy $dE$. The majority of analyses based on this type of event counting deduce $\\alpha\\approx 1.7$ \\citep{Lin84, Shimizu95, AP02}, a value smaller than the critical 2. These results may, however, be misleading. For example, \\citet{Parnell04} demonstrated that one can obtain $\\alpha$ ranging from 1.5 to 2.6 for the same data set using different but still reasonable sets of assumptions for the analyses. \\begin{figure} \\includegraphics[width=8.1cm,angle=0]{0911fig1.eps} \\caption{ EIT 195 \\AA ~images of the observed active region at two times, showing the position of the SUMER slit, indicated by the vertical line.} \\label{fig1} \\end{figure} Here we take an alternative approach and model ultraviolet (UV) radiances observed by the Solar Ultraviolet Measurements of Emitted Radiation \\citep[SUMER;][]{Wetal95, Wetal97} in an active region corona, assuming that the radiance fluctuations and the nearly constant `background' emission are caused by small-scale stochastic flaring \\citep{PS04b, PS06}. The model has been applied successfully to UV radiance fluctuations in the quiet Sun \\citep{PS06}. The method compares light curves generated assuming random flaring with a power law frequency distribution to the light curves of an observed emission line. It has the advantage that it takes into account without bias weak, blended micro- and nanoflares that produce a nearly continuous background. Here we apply this technique to off-limb time series recorded by SUMER. The three lines modelled, \\FeXIX~$\\lambda$\\,1118.07 (6.3~MK), \\CaXIII~$\\lambda$\\,1133.76 (2.2~MK) and \\SiIII~$\\lambda$\\,1113.23 (0.06~MK), cover two decades of formation temperature from the lower transition region to the hotter gas in the corona. The analysis described here uses Artificial Neural Networks (ANNs) to find the optimum match to the three parameters of the model. The main advantage of this method over previous analyses based on the radiance distribution function \\citep{PS06, SISP07} is that we are able to obtain quantitative values for all parameters, including $\\alpha$. Another advantage of the ANN method is that it concentrates on the number and shape of the emission peaks along the light curves with little weight on the low radiance pixels, which was a problem with the \\citet{SISP07} analysis. ", "conclusions": "\\label{results} \\subsection{Results} In the present work, PNN is used as a tool to extract the three flare model parameters required to reproduce the SUMER light curves. All 35 \\FeXIX\\ and \\CaXIII, and 11 \\SiIII\\ SUMER light curves from the three days of observations were fed individually into the neural network and the parameters were obtained for each light curve separately. The final PNN outputs are shown in Table \\ref{tab1}. The bold numbers are the statistically maximum occurrence for each parameter. For example for \\FeXIX, $\\alpha=2.8$ is found in more than 70\\% of the light curves. The minimum and maximum values, given on the left and right, indicate the scatter in the light curve parameters. \\begin{table} \\centering \\caption{The SUMER spectral lines and the parameter values given by PNN.} \\label{tab1} \\smallskip \\begin{threeparttable} \\begin{tabular}{c c c c} \\hline\\hline\\\\ SUMER spectra & \\multicolumn{3}{c}{PNN outputs\\tnote{1}}\\\\ lines & $\\alpha$ & $\\tau$ & $p_f$ \\\\ \\hline \\hline \\\\ \\FeXIX & 2.5 {\\bf 2.8} 3.1 & 9 {\\bf 14} 20 & 0.1 {\\bf 0.2} 0.7 \\\\ \\CaXIII & 2.6 {\\bf 2.8} 3.0 & 41 {\\bf 45} 45 & 0.8 {\\bf 0.9} 0.9 \\\\ \\SiIII & 2.4 {\\bf 2.6} 2.8 & 5 {\\bf 9} 12 & 0.2 {\\bf 0.3} 0.9 \\\\ \\hline \\end{tabular} \\begin{tablenotes} \\item[1]The most frequent values are given in bold, and the minimum and maximum on the left and right. $\\tau$ is per exposure time (90 s) and $p_f$ is per exposure time per 5\\arcsec$\\times$4\\arcsec\\ spatial element. \\end{tablenotes} \\end{threeparttable} \\end{table} In each line there is 20\\% scatter in $\\alpha$, and 50\\% scatter in $\\tau$. The range of $p_f$ values for \\FeXIX\\ and \\SiIII\\ is much broader, suggesting that events producing emission in these temperature ranges do not have the same rate everywhere but are seen in irregular bursts. We also note that the value of $\\tau p_f$ is roughly the same for both \\FeXIX\\ and \\SiIII, as suggested by their shape parameter (Fig.~\\ref{timeseries}). The \\CaXIII\\ light curves are all matched with a high value of $p_f$, consistent with the idea that the 1~MK active region corona requires almost continuous flaring. The four times higher rate for \\CaXIII\\ than \\FeXIX\\ suggests that most of the \\CaXIII\\ emission is produced by heating events below the \\FeXIX\\ formation temperature (6.6~MK). Example light curves obtained using these parameters are compared with the observed ones in Fig.~\\ref{lightcurves}. Both the \\SiIII\\ and \\CaXIII\\ simulations look remarkably similar to their observed light curves. The background radiance of the \\FeXIX\\ light curve is about a factor of 2 too low. The \\FeXIX\\ light curves had a $p_f$ ranging from 0.7 to 0.1, so we suspect that in this case the $p_f$ value is slightly too low. Also for \\FeXIX, the ratio $\\tau_r/\\tau_d$ deduced from the data is smaller than the fixed value 0.5 used here. This may influence the accuracy of the method. The sensitivity of the PNN output depends on the training set. During the training session, the network must see all possible patterns that it is supposed to classify in the testing session. With 500 simulated light curves in the training set, PNN was not able to converge for several of the SUMER light curves. When we increased the number of simulated light curves to 6930, we were able to obtain unique parameters for all observed light curves. \\begin{figure} \\includegraphics[width=7.5 cm]{0911fig6.eps} \\caption{Samples of the radiance time series: left panel: SUMER data, and right panel: simulation data obtained with the parameters given in Table 1. } \\label{lightcurves} \\end{figure} \\subsection{Conclusions} The concept that the solar corona may be heated by numerous, randomly distributed, small flare-like events called nanoflares is considered by comparing simulated and observed emission line light curves. The difference between this and previous methods is the fully automated modelling of the light curve structure. There is no human decision required for background/event cut-off levels or best fit parameters. The result is power law flare energy frequency exponents greater than 2.5 for all three emission lines considered, \\SiIII, \\CaXIII\\ and \\FeXIX. This is consistent with the corona being heated mainly by nanoflares, and demonstrates the importance of nanoflare 'background' emission in determining the power law exponents. The parameter with highest uncertainty or largest scatter is the flare rate, especially for the lines formed at transition region and hot flare temperatures. Coronal plasma at these temperatures is produced sporadically and is associated with more specific coronal and chromospheric loop structures than the general active region corona, so the scatter is to be expected. The next step will be to determine the actual flare energies producing the nanoflare emission. This is a much more complicated exercise because the modelled light curves are observed in the corona which may be heated by events occurring lower in the atmosphere \\citep{Aschwanden08}, so that it requires a model for the energy transfer to the observation position." }, "0808/0808.3578_arXiv.txt": { "abstract": "We introduce a new code for computing time-dependent continuum radiative transfer and non-equilibrium ionization states in static density fields with periodic boundaries. Our code solves the moments of the radiative transfer equation, closed by an Eddingtion tensor computed using a long characteristics method. We show that pure (i.e., not source-centered) short characteristics and the optically-thin approximation are inappropriate for computing Eddington factors for the problem of cosmological reionization. We evolve the non-equilibrium ionization field via an efficient and accurate (errors $<1\\%$) technique that switches between fully implicit or explicit finite-differencing depending on whether the local timescales are long or short compared to the timestep. We tailor our code for the problem of cosmological reionization. In tests, the code conserves photons, accurately treats cosmological effects, and reproduces analytic Str\\\"omgren sphere solutions. Its chief weakness is that the computation time for the long characteristics calculation scales relatively poorly compared to other techniques ($\\tlc \\propto N_{\\rm{cells}}^{\\sim1.5}$); however, we mitigate this by only recomputing the Eddington tensor when the radiation field changes substantially. Our technique makes almost no physical approximations, so it provides a way to benchmark faster but more approximate techniques. It can readily be extended to evolve multiple frequencies, though we do not do so here. Finally, we note that our method is generally applicable to any problem involving the transfer of continuum radiation through a periodic volume. ", "introduction": "\\label{sec:intro} The epoch of reionization is the current frontier in understanding how galaxies form and evolve over cosmic time. After the Universe cooled sufficiently to recombine hydrogen atoms at redshift $z\\approx 1088$~\\citep{spe07}, the Universe was fully neutral. Gravity grew ever-denser structures that, at $z\\sim 30-50$, were able to collapse into stars and/or black holes. The radiation emitted from these first objects then began to re-ionize hydrogen. By $z\\sim 6$, hydrogen reionization appears to be complete~\\citep{fan07}, and the diffuse intergalactic medium (IGM) has a neutral fraction of $\\sim 10^{-4}$. Understanding this transition epoch is central to understanding the origin of galaxies and the evolution of the IGM. It is a major science driver for a host of upcoming international telescope facilities, such as the {\\it James Webb Space Telescope} and the {\\it Atacama Large Millimeter Array}. Reionization involves a complex interplay between nonlinear growth of structure, radiative cooling, star/black hole formation, chemical enrichment, and photon transport. Numerical simulations are required to accurately model these highly nonlinear processes. However, the large dynamic range and complex physics involved make this an extraordinarily challenging computational problem. To obtain a full picture of reionization in the context of currently-favored hierarchical structure formation models, it is imperative that simulations include processes of star formation, galaxy formation, and IGM evolution, along with feedback processes that connect all three. Cosmological hydrodynamic simulations accounting for these processes are now achieving maturity, thanks to improving algorithms and computing power. However, the inclusion of radiation transport complicates matters immensely. A cosmological radiative hydrodynamics code that can accurately evolve a representative volume with sufficient dynamic range to study how galaxies reionize the Universe would be a major development towards understanding reionization. In this paper we provide a step towards that end by introducing a new accurate moment-based method for calculating radiative transfer (RT) in a cosmological context. Time-dependent radiative transfer is one of the most difficult components to treat in any theoretical study of the reionization epoch owing to the problem's well-known high dimensionality. Consequently, over the past decade, a number of approximate treatments have emerged that seek to render it more tractable through well-motivated physical approximations. The most flexible methods are the fully analytic treatments~\\citep[for example,][]{mad99,wyi03,fu04,ili05,kra06}. These generally involve assuming values for quantities such as the gas clumping factor and the recombination rate that are averaged over all space or, in the case of the excursion-set formalism~\\citep{fu04}, over the volume of an ionized region. In exchange, they readily allow for broad surveys of parameter space to be performed. The next step in the direction of a full solution is taken by the semi-numerical methods~\\citep{cia00,mes07,gei08,cho08}, which combine numerically-generated density fields with analytic treatments for radiative transfer using techniques such as the excursion-set formalism in order to account more realistically for source bias and the effects of inhomogeneous density fields. These treatments offer a dramatic increase in realism over purely analytic calculations at modest additional computational cost. However, they have some difficulty accounting fully for the consequences of inhomogeneous density fields such as shadowing and the tendency for low-density regions to have a lower neutral fraction during the later stages of reionization~\\citep{cho08}. These problems arise from the need of semi-numerical models to make assumptions regarding the shape of the ionized regions surrounding individual sources and the nontrivial relationship between dark matter and gas densities in the nonlinear regime. Some of these difficulties are avoided in models that actually solve the radiative transfer equation on numerically-generated density fields but without fully accounting for radiative feedback on the sources (\\citealt{cia01,sok01,mel06,mcq07,ili07a}; see also \\citealt{ili06} for a very useful comparison of a number of techniques). Nonetheless, obtaining realistic baryonic density and emissivity fields in such contexts still presents considerable challenges~\\citep{mcq07}. Additionally, while parametrized treatments for radiative feedback have been introduced in such models in order to study, for example, whether the photoevaporation of minihaloes extends the epoch of reionization~\\citep{ili07a,mcq07}, the simplified nature of these studies leaves their results open to question~\\citep{mes07}. Hence, while each of these methods has yielded an abundance of insight into reionization and warrant continued development, the need is emerging for a complete solution to the radiative transfer equation that is merged self-consistently with hydrodynamical calculations. A few fully radiative-hydrodynamic codes have been used to study the reionization epoch~\\citep{gne01,cen02,rij06}. Unfortunately, the techniques that these codes have introduced do not yet enjoy widespread use owing to their high computational expense, even though such studies are crucial for tuning the assumptions employed in more simplified treatments and verifying their conclusions. Moreover, despite the enormous gain in realism that these codes have afforded the reionization community, they too involve some physical approximations such as the use of the ``optically thin variable Eddington tensor\" approximation~\\citep{gne01} and a reduced~\\citep{gne01} or an infinite speed of light~\\citep{cen02,rij06}. In this work, we present a moment method solution to the radiative transfer equation~\\citep{aue70} and test it on static density fields. Our technique is highly flexible, involves a minimum of physical approximations, and can readily be combined with existing hydrodynamical calculations. It is similar to the method presented by~\\citet{sto92}, but with several differences. First, we optimize our code only for cubical simulation volumes with periodic boundaries, as this is typical of cosmological simulations. Second, we derive our Eddington tensors from a long characteristics (LC) calculation in order to minimize artifacts owing to poor angular and spatial resolution. Finally, we include a treatment for nonequilibrium ionizations and account for the cosmological terms in the radiative transfer equation. In a follow-up paper, we will present its implementation within a cosmological galaxy formation code. We begin in Section~\\ref{sec:rt} by casting the radiative transfer equation into the form in which we solve it and summarizing our numerical method. In Section~\\ref{sec:fedd}, we compare the performance of long characteristics versus two other time-independent radiative transfer techniques in order to select a method for deriving the Eddington tensor, which we need in order to close our moment hierarchy. After demonstrating that long characteristics introduces the fewest unphysical artifacts, we optimize it for computing reionization using a suite of realistic albeit low-resolution integrations. In Section~\\ref{sec:nlte}, we discuss our technique for evolving the nonequilibrium ionization field. In Section~\\ref{sec:full_alg}, we summarize our iterative scheme for weaving these ingredients into a self-consistent calculation. In Section~\\ref{sec:tests}, we subject our code to a number of standard tests. Finally, we summarize our method and results in Section~\\ref{sec:summary}. ", "conclusions": "\\label{sec:summary} In this paper, we introduced a method that accurately and efficiently computes continuum radiative transfer in static density fields. The code uses a moment-based approach to solve the equation of comoving radiative transfer, with the Eddington tensors obtained using a long characteristics method. We compared several techniques for computing the Eddington tensors that are needed to close the moment hierarchy and demonstrated that, of the three methods that we investigated, only the method of long characteristics has the ability to compute highly inhomogeneous radiation fields without introducing numerical artifacts. We found through direct measurement that the computation times for our long characteristics and moments modules scale with the number of computational cells $N_{\\rm{cells}}$ as $N_{\\rm{cells}}^{1.5}$ and $N_{\\rm{cells}}^{1.0}$, respectively. Next, we introduced a hybrid method for computing the evolution of nonequilibrium ionization fields and demonstrated that it is accurate to 1\\% throughout our computational domain. We combined this with our radiative transfer code via an efficient iterative algorithm. The final code is regulated by a number of parameters, and we characterized how these parameters impact computation time and accuracy using a suite of low-resolution convergence tests. We subjected our method to a number of standard problems in continuum radiative transfer. First, we verified that our code accurately computes the growth of an \\hii\\ region about a source in the classical (static) case as well as in the case of an expanding medium. We found that, in both cases, the radius of the resulting ionized region agrees with analytic expectations to within the computation's spatial resolution at all times. Next, we tested whether our code is able to produce shadows behind opaque regions. In agreement with previous work, we found that the moment method introduces a small diffusion into the shadowed region~\\citep{hay03}. Nevertheless, we found that the strength of the radiation field in the shadowed region was up to only 1\\% of the value in the unshadowed region. Finally, we computed the reionization of an expanding cosmological volume and found qualitative agreement with other work in the literature. Our code currently accounts for radiative and collisional ionization of hydrogen and helium as well as radiative recombination. It does not account for recombination radiation. Additionally, it does not follow the evolution of the temperature field, hence it does not account for photoionization suppression of star formation in low-mass halos, photoionization heating, recombination cooling, or shock formation. In the future we plan to expand on our code in several ways. First, we have found that the computation time increases dramatically as the universe becomes optically thin because the LC line integrals traverse more cells before terminating either because they arrive at the source or because they reach $\\taumax$. However, in this regime, the optically-thin approximation, which is roughly ten times faster than LC, becomes increasingly valid. For this reason, we plan to study how to transition smoothly from LC to the optically thin approximation without introducing accuracy errors. Second, we will generalize our method to multifrequency radiative transfer, which is necessary for studying, for example, ionization front hardening or \\heii\\ reionization. Finally, we plan to merge our radiative transfer scheme with our version of the cosmological galaxy formation code {\\sc GADGET-2}." }, "0808/0808.1261_arXiv.txt": { "abstract": "Due to quantum fluctuations, spacetime is foamy on small scales. The degree of foaminess is found to be consistent with the holographic principle. One way to detect spacetime foam is to look for halos in the images of distant quasars. Applying the holographic foam model to cosmology we \"predict\" that the cosmic energy density takes on the critical value; and basing only on existing archived data on active galactic nuclei from the Hubble Space Telescope, we also \"predict\" the existence of dark energy which, we argue, is composed of an enormous number of inert ``particles\" of extremely long wavelength. We speculate that these ``particles\" obey infinite statistics. ", "introduction": "Like everything else, spacetime is conceivably subject to quantum fluctuations. So we expect that spacetime, probed at a small enough scale, will appear complicated --- something akin in complexity to a turbulent froth that John Wheeler has dubbed \"quantum foam,\" also known as \"spacetime foam.\"\\footnote{In the gravitational context, the phenomenon of turbulence is indeed intimately related to the properties of spacetime foam. See Ref. \\cite{Jejjala2008}.} But how large are the fluctuations in the fabric of spacetime? To quantify the problem, let us recall that, if spacetime indeed undergoes quantum fluctuations, there will be an intrinsic limitation to the accuracy with which one can measure a distance, for that distance fluctuates. Denoting the fluctuation of a distance $l$ by $\\delta l$, on general grounds, we expect $\\delta l \\gtrsim l^{1 - \\alpha} l_P^{\\alpha}$, where $l_P = \\sqrt{\\hbar G/c^3}$ is the Planck length, the characteristic length scale in quantum gravity, and we have denoted the Planck constant, gravitational constant and the speed of light by $ \\hbar$, $G$ and $c$ respectively. The parameter $\\alpha \\sim 1$ specifies the different spacetime foam models. In this talk we will concentrate on the model corresponding to $\\alpha = 2/3$, which has come to be known as the holographic model \\cite{wigner,Karol}, so called because it is found to be consistent with the holographic principle \\cite{tHooft,susskind}, according to which, the information content inside any three dimensional region of space can be encoded on the two dimensional surface around the region, like a hologram. For comparison, we will also consider the random-walk model \\cite{Amelino1999} corresponding to $\\alpha = 1/2$. Contents of this talk: Applying nothing more than quantum mechanics and some rudimentary black hole physics, we derive the holographic model of spacetime foam. Applying the holographic model to cosmology, we \"predict\" that the cosmic energy density takes on the critical value (i.e., the fractional density parameter of the universe $\\Omega \\cong 1$), consistent with observation. Then aided by some archived data on quasar or AGN from the Hubble Space Telescope, we are led to conclude that dark energy exists. Furthermore we are naturally led to speculate that the constituents of dark energy, unlike ordinary matter, obey an exotic statistics known as infinite statistics in which all representations of the particle permutation group can occur. ", "conclusions": "" }, "0808/0808.0267_arXiv.txt": { "abstract": "The primary scientific goal of the GRIPS mission is to revolutionize our understanding of the early universe using $\\gamma$-ray bursts. We propose a new generation gamma-ray observatory capable of unprecedented spectroscopy over a wide range of $\\gamma$-ray energies (200 keV--50 MeV) and of polarimetry (200--1000 keV). Secondary goals achievable by this mission include direct measurements of supernova interiors through $\\gamma$-rays from radioactive decays, nuclear astrophysics with massive stars and novae, and studies of particle acceleration near compact stars, interstellar shocks, and clusters of galaxies. ", "introduction": "Gamma-ray bursts (GRB) are the most luminous sources in the sky, and thus act as signposts throughout the Universe. The long-duration sub-group is produced by the explosion of massive stars, while short-duration GRBs likely originate during the merging of compact objects. Both types are intense neutrino sources, and being stellar sized objects at cosmological scales, they connect different branches of research and thus have a broad impact on present-day astrophysics. Identifying objects at redshift \\gax 6.5 has become one of the main goals of modern observational cosmology, but turned out to be difficult. GRBs offer a promising opportunity to identify high-$z$ objects, and moreover even allow us to investigate the host galaxies at these redshifts. GRBs are a factor 10$^{5-7}$ brighter than quasars during the first hour after explosion, and a favourable relativistic k-correction implies that they do not get fainter beyond $z$$\\sim$3. Yet, present and near-future ground- and space-based sensitivity limits the measurement of redshifts at $z$$\\sim$13 (as $H$-band drop-outs), because GRB afterglows above 2.5 $\\mu$m are too faint by many magnitudes % for 8--10\\,m telescopes. % Thus, a completely different strategy is needed to step beyond redshift 13 to measure when the first stars formed. Fortunately, nuclear physics offers such a new strategy. Similar to X-ray and optical absorption lines due to transitions between electronic levels, resonant absorption processes in the nuclei exist which leave narrow absorption lines in the $\\gamma$-ray range. The most prominent and astrophysically relevant are the nuclear excitation and Pygmy resonances (element-specific narrow lines between 5--9 MeV), the Giant Dipole resonance (GDR; proton versus neutron fluid oscillations; $\\sim$ 25 MeV; two nucleons and more) and the Delta-resonance (individual-nucleon excitations, 325 MeV; all nucleons, including H!). Such resonant absorption only depends on the presence of the nucleonic species, and not on ionization state and isotope ratio. They imprint well-defined spectral features in the otherwise featureless continuum spectra of GRBs (and other sources). This is completely new territory (Iyudin \\etal\\ 2005), but with the great promise to measure redshifts directly from the gamma-ray spectrum, i.e. without the need for optical/NIR identification! Technically, this new strategy requires sensitive spectroscopy in the 0.2--50 MeV band. The detection of GRBs requires a large field of view. Therefore, the logical detection principle is a Compton telescope. In addition, such detectors can be tailored to have a high polarisation sensitivity. Polarimetry is the last property of high-energy electromagnetic radiation which has not been utilized in its full extent, and promises to uniquely determine the emission processes in GRBs, as well as many other astrophysical sources. With its large field of view, such a detector will not only scan 80\\% of the sky within one satellite orbital period of 96 min., but also provide enormous grasp for measuring the diffuse emission of nucleosynthesis products and cosmic-ray acceleration. ", "conclusions": "" }, "0808/0808.2180_arXiv.txt": { "abstract": "We have determined the distance to NGC~4258 using observations made with the Hubble Space Telescope (HST) and the Wide Field, Advanced Camera for Surveys (ACS/WFC). We apply a modified technique that fully accounts for metallicity effects on the use of the luminosity of the tip of the red giant branch (TRGB) to determine one of the most precise TRGB distance moduli to date: $\\mu(TRGB) = 29.28 \\pm 0.04$ (random) $\\pm 0.12$ (systematic) mag ($7.18 \\pm 0.13 \\pm 0.40$ Mpc). We discuss this distance modulus with respect to other recent applications of the TRGB method to NGC~4258, and with several other techniques (Cepheids and masers) that are equally competitive in their precision, but different in their systematics. ", "introduction": "This is the first in a short series of papers using a refined methodology for determining distances using the discontinuity in the I-band magnitude of the red giant branch luminosity function as a standard candle (Lee et al. 1993), the so-called TRGB (tip of the red giant branch) method. Here we apply a new methodology in correcting for the now well understood and precisely calibrated metallicity effects on the TRGB magnitude (see Section 4.1 and Madore et al. 2008). Our first target is the spiral galaxy NGC~4258. It is nearby, and therefore very highly resolved, not only into its bright, high-mass Population I disk stars, but also into its fainter, but still accessible, low-mass Population II halo stars. NGC~4258 contains many known Cepheids that have been discovered and used as distance indicators in multiple observing campaigns using HST. Its halo has been resolved and studied on equally as many occasions, revealing a broad, richly populated giant branch for TRGB distance determination. The uniqueness of NGC~4258 lies at its center, where a Keplerian-rotating disk of water masers has proper motions and radial velocities that can be cross-compared and modeled with essentially one additional free parameter: the distance. As such, the independently calibrated Population~I (Cepheid) and Population~II (TRGB) distance scales both converge on and cross at NGC~4258, where they can be compared to that from simple geometry (maser method). No other galaxy provides such an environment for testing the distance scale. That said, it must also be emphasized that NGC~4258 is still only one object, and its uniqueness means that there is no independent check on the maser distance methodology itself, its random errors, or its systematics. Without prejudice as to which (if any) of the three distance determination methods discussed here is better (understood or calibrated) at this point, we now proceed to present a new and improved determination of the TRGB distance using HST ACS/WFC data from one of our approved and scheduled programs, and archival WFPC2 data as a consistency check. We compare these results with previous TRGB results, and with the other past and published methods. ", "conclusions": "" }, "0808/0808.0185_arXiv.txt": { "abstract": "We have acquired near-infrared spectra of Kuiper belt objects 2003 UZ117, 2005 CB79 and 2004 SB60 with NIRC on the Keck I Telescope. These objects are dynamically close to the core of the 2003 EL61 collisional family and were suggested to be potential fragments of this collision by \\citet{darin}. We find that the spectra of 2003 UZ117 and 2005 CB79 both show the characteristic strong water ice absorption features seen exclusively on 2003 EL61, its largest satellite, and the six other known collisional fragments. In contrast, we find that the near infrared spectrum of 2004 SB60 is essentially featureless with a fraction of water ice of less than 5\\%. We discuss the implications of the discovery of these additional family members for understanding the formation and evolution of this collisional family in the outer solar system. ", "introduction": "The only known collisional family in the Kuiper belt was detected because of the unique spectral properties the family members (Brown et al. 2007). The near-infrared spectra of all other non-volatile rich Kuiper belt objects (KBOs) lie on a continuum between those whose spectra contain moderate amounts of water ice absorptions to those whose spectra are essentially featureless \\citep{2007Natur.446..294B, kris08}. In contrast, the near-infrared spectra of 2003 EL61, its largest satellite, and five other small KBOs resemble laboratory spectra of pure crystalline water ice \\citep{1999ApJ...519L.101B,kris06,2007ApJ...655.1172T,2007A&A...466.1185M,2007Natur.446..294B, kris08}. Remarkably, these objects are also relatively clustered in orbital element space. The largest object, 2003 EL61, had been previously suggested to have experienced a massive collision that imparted its fast (4-hour) rotation, stripped off most of its icy mantle leaving it with a density close to rock (2.7 g/cc), and formed its two satellites \\citep{2006ApJ...639.1238R, 2006ApJ...639L..43B, 2008AJ....135.1749L}. \\citet{2007Natur.446..294B} concluded that the four extremely water ice rich KBOs and the large satellite of 2003 EL61 were in fact fragments of the icy mantle of the proto-2003 EL61 that had been ejected during a massive collision. While collisional families in the asteroid belt can be identified by their dynamical clustering alone, families in the Kuiper belt are much harder to identify because the collisional ejection velocities can be a significant fraction of an object's orbital velocity. Collisional fragments can therefore be spread out over a wide range of orbital element space with the fastest ejected fragments having significantly different orbits from the family core. The 2003 EL61 family members were all identified on the basis of their unique surface properties: they are the only known objects in the Kuiper belt with extremely pure water ice spectra and neutral visible colors. It is only because of this unique surface signature that detection of the 2003 EL61 family members was possible. In order to determine if other known KBOs could be fragments of the 2003 EL61 collision, \\citet{darin} integrated the orbits of 131 high inclination KBOs to determine their proper orbital elements and minimum ejection velocities away from the 2003 EL61 family core. They determined the minimum ejection velocity a KBO must have had in order to reach its present orbit from the modeled location of the family forming impact. In Figure 1 we show object H-magnitude vs. minimum ejection velocity (from \\citet{darin}) for the known KBOs closest to the family core. For KBOs in known resonances, we plot the ejection velocity accounting for resonance diffusion \\citep{darin}. A large fraction of these objects have never been observed to determine if they have IR spectral signatures consistent with the 2003 EL61 family members. It is important to note that though all of the known 2003 EL61 family members have neutral visible colors, visible color or visible spectroscopy alone \\citep{PinAlons2008} is not sufficient for family member identification. There are many KBOs with neutral visible colors that do not also have strong water ice absorptions and are not members of the 2003 EL61 collisional family. Therefore, while visible colors can be used to rule out objects as potential family members, near infrared spectroscopy to determine water ice absorption depths is necessary for definitive family member identification. Identifying additional family members and characterizing the extent of the spread of fragments throughout the Kuiper belt may provide insight into the physics of this giant impact event in the outer solar system. In this letter we present near-infrared spectra obtained with the Keck I telescope of KBOs 2003 UZ117, 2005 CB79 and 2004 SB60, objects that are located 67, 97, and 221 m/s away from the modeled 2003 EL61 family core respectively (Fig 1). 2003 UZ117 and 2005 CB79 were suggested to be the most likely additional family candidates by \\citet{darin}. ", "conclusions": "2003 UZ117 and 2005 CB79 contain high fractions of pure water ice on their surfaces and appear dynamically related to the other members of the 2003 EL61 collisional family. We therefore conclude that they are members of this collisional family. Thus far, all objects within 130 m/s of the modeled collision center \\citep{darin} have been shown to have the same unique strong water ice spectral signatures. 2004 SB60, at 221 m/s from the family core, has an essentially featureless spectrum inconsistent with 2003 EL61 and its family members but comparable to many other small KBOs. The flat spectrum we observed is consistent with infrared photometric observations of 2004 SB60 with HST by \\citet{2007DPS....39.3906S}. Infrared spectral or spectrophotometric observations of more potential fragments with minimum ejection velocities from $\\sim$150 to $\\sim$300 m/s could help constrain the extent to which fragments were scattered throughout the Kuiper belt and could tell us about the physics of the giant impact itself. In addition to the distribution of fragments, any collisional model would also need to explain the the presence of the two satellites of 2003 EL61. Though it does not have the spectral signature of the fragments of the 2003 EL61 collision, 2004 SB60 is an interesting object in its own right. \\citet{2008Icar..194..758N} have found this object to be one of the few high inclination small binary objects in the Kuiper belt. It is worth noting the possibility that not all fragments of the 2003 EL61 collision necessarily have the unique strong water ice signature. Fragments from different locations in the initial parent body may have had different initial compositions. However, without this unique spectral signature, identification of an object as a family member is currently impossible. Another object near the center of the 2003 EL61 family is the third largest KBO, 2005 FY9. 2005 FY9 has a velocity closer to the center of the collision (150 m/s) than one of the known fragments but has a spectrum dominated by methane, not water ice absorptions. We find it intriguing that two of the largest KBOs are so close to each other in orbital element space but can think of no reason why they should be physically related. The spectral signature of extremely pure water ice found on 2003 EL61 and its fragments and nowhere else in the Kuiper belt is the only reason definitive identification of the collisonal family was possible. A collision with a non differentiated parent body would likely not produce such a distinct spectral signature in its fragments. Future large scale surveys of the Kuiper belt such as PanSTARRS and LSST are expected to increase the numbers of known KBOs by over an order of magnitude. With higher numbers of objects, overdensities in certain regions may be revealed and families without unique spectral signatures may be able to be identified by their dynamics alone." }, "0808/0808.3904_arXiv.txt": { "abstract": "{} {We present an accurate characterisation of the high-resolution X-ray spectrum of the Narrow Line Seyfert 1 galaxy Arakelian 564 and put it in to context with other objects of its type by making a detailed comparison of their spectra.} {The data are taken from 5 observations with the \\textit{XMM-Newton} Reflection Grating Spectrometer and fitted with various spectral models.} {The best fit to the data identifies five significant emission lines at 18.9, 22.1, 24.7, 29.0 and 33.5\\AA~due O\\,VIII Ly$\\alpha$, O\\,VII(f), N\\,VII Ly$\\alpha$, N\\,VI(i) and C\\,VI Ly$\\alpha$ respectively. These have an RMS velocity of $\\sim1100$\\,km\\,s$^{-1}$ and a flow velocity of $\\sim-600$\\,km\\,s$^{-1}$, except for the O\\,VII(f) emission line, which has a flow velocity consistent with zero. Two separate emitting regions are identified. Three separate phases of photoionized, X-ray absorbing gas are included in the fit with ionization parameters log\\,$\\xi=-0.86$, 0.87, 2.56 and column densities $N_H=0.89, 2.41, 6.03\\times10^{20}$\\,cm$^{-2}$ respectively. All three phases show this to be an unusually low velocity outflow ($-10\\pm100$\\,km\\,s$^{-1}$) for a narrow line Seyfert 1. We present the hypothesis that the BLR is the source of the NLR and warm absorber, and examine optical and UV images from the \\textit{XMM-Newton} Optical Monitor to relate our findings to the characteristics of the host galaxy.} {} ", "introduction": "\\label{564intro} Narrow Line Seyfert 1 galaxies (NLS1s) were initially classified by \\cite{osterbrock85} and are defined as having H$\\beta$ FWHM\\,$\\leq$\\,2000\\,km\\,s$^{-1}$. Their X-ray spectra often display more rapid variability (e.g.~\\nocite{boller96}Boller et al. 1996;~\\nocite{laor94}Laor et al. 1994) and are steeper in the soft X-ray band than those of `normal' broad line Seyfert 1s (e.g.~\\nocite{leighly99}Leighly 1999). From the \\textit{ROSAT} All Sky Survey it was found that approximately half of the sources in soft X-ray selected samples of AGN are NLS1s (\\nocite{grupe96}Grupe 1996;~\\nocite{hasinger97}Hasinger 1997). They are believed to be `high-state' active galaxies, with low black hole masses for their luminosity, and so with high accretion rates relative to Eddington (\\nocite{pounds95}Pounds et al. 1995;~\\nocite{boroson02}Boroson 2002). \\newline Observations indicate that the majority of Seyfert 1s display evidence for `warm absorbers' (\\nocite{blustin05}Blustin et al. 2005), i.e. clouds of photoionized X-ray absorbing gas in our line of sight (e.g.~\\nocite{komossa00}Komossa 2000), yet still very little is known of the absorbers origin, location or structure. The warm absorber is often characterised by its ionization parameter, $\\xi$. This is a measure of the ionization state of the material and is defined as \\begin{equation} \\xi=\\frac{L}{nr^2} \\label{564_xi_eq} \\end{equation} where $L$ is the source luminosity (in erg\\,s$^{-1}$), $n$ is the gas number density (in cm$^{-3} $) and $r$ is the distance of the absorber from the central engine (in cm) (\\nocite{tarter69}Tarter et al. 1969). \\newline Arakelian 564 is the brightest known NLS1 in the 2--10\\,keV range (L$_{2-10\\,keV}=2.4\\times10^{43}\\,$erg\\,s$^{-1}$, \\nocite{turner01} Turner et al. 2001). This, coupled with its close proximity, z=0.02468 (\\nocite{huchra99}Huchra et al. 1999), and relatively low Galactic column density, N$_H=5.34\\times10^{20}$cm$^{-2}$ (\\nocite{kalberla05}Kalberla et al. 2005), make it a very interesting target to investigate. This object has been studied across all wavebands (e.g.~\\nocite{shemmer01}Shemmer et al. 2001;~\\nocite{romano04}Romano et al. 2004). Here we have merged data from 5 observations with the \\textit{XMM-Newton} Reflection Grating Spectrometer (RGS), allowing us to explore, in unprecedented detail, the high-resolution X-ray spectrum of this object. \\newline \\citet{matsumoto04} report on a 50\\,ks \\textit{Chandra} HETGS observation. They fit the hard X-ray spectrum with a power law of photon index of $2.56\\pm0.06$ and add a blackbody, of temperature $0.124\\pm0.003$\\,keV, to fit the soft excess. They detect an edge-like feature at 0.712\\,keV, and two phases of absorption by photoionized gas (phase 1: log\\,$\\xi\\sim1$, N$_{H}\\sim10^{21}$\\,cm$^{-2}$; phase 2: log\\,$\\xi\\sim2$, N$_{H}\\sim10^{21}$\\,cm$^{-2}$).~\\citet{vignali04} have analysed two \\textit{XMM-Newton} observations from June 2000 and June 2001. They fit the spectrum with a steep power law ($\\Gamma=2.50-2.55$) plus a soft blackbody component (kT$\\sim$140\\,--\\,150\\,eV). They also identify an edge-like feature in the EPIC data at $\\sim0.73$\\,keV, which they interpret as an O\\,VII K absorption edge, and find evidence for a broad iron emission line, at $\\sim1$\\,keV, in one observation. ~\\citet{crenshaw02} have analysed the UV spectrum using data from the STIS onboard the \\textit{Hubble Space Telescope}. They identify UV absorption lines due to Ly$\\alpha$, N\\,V, C\\,IV, Si\\,IV and Si\\,III centred at a radial velocity of $-190$\\,km\\,s$^{-1}$ relative to the systemic velocity. They identify this with a dusty lukewarm absorber with column density N$_{H}=1.62\\times10^{21}$\\,cm$^{-2}$ and ionization U=0.033 (this ionization is equivalent to log\\,$\\xi\\simeq1.2$, converted using the method of \\nocite{george98}George et al. 1998). \\newline \\citet{papadakis06} analysed the most recent (2005), and longest ($\\sim100$\\,ks), observation with \\textit{XMM-Newton}, focusing on the EPIC data. They find that the soft excess cannot be parametrized either by a multiple power law or by a power law plus a single blackbody, in contrast to previous findings. They fit the MOS and PN spectra with either a power law of slope $2.43\\pm0.03$ and two blackbodies, with kT$\\sim$0.15 and $\\sim$0.07\\,keV, or with a relativistically blurred photoionized disk reflection model. They also find evidence for two phases of photoionized X-ray absorbing gas with ionization parameters log\\,$\\xi\\sim1$ and $\\sim2$ and column densities N$_{H}\\sim2$ and $5\\times 10^{20}$\\,cm$^{-2}$. \\newline The same 2005 data were also analysed by~\\citet{dewangan07}. They concentrate their analysis on the EPIC data, but also study the RGS spectrum and find two warm absorber phases with ionization parameters log\\,$\\xi\\sim2$ and $<0.3$, column densities $N_H\\sim4\\times10^{20}$ and $2\\times10^{20}$cm$^{-2}$ and flow velocities $v\\sim-300$ and $-1000$\\,km\\,s$^{-1}$ respectively. \\newline Here we investigate the combined RGS spectrum of Arakelian 564. Section~\\ref{obs_data_red} gives information on the individual observations and describes our data reduction process. The spectral fitting and best fit results are presented in Section~\\ref{analysis}. A full discussion of these results and comparison with other observations is given in Section~\\ref{discussion}. ", "conclusions": "We present a detailed analysis of the photoionized, X-ray absorbing and emitting gas present in the NLS1 galaxy Arakelian 564. The high-resolution X-ray spectra are from five observations with \\textit{XMM-Newton}'s Reflection Grating Spectrometer. \\newline The absorption profile is fitted with three phases of X-ray absorbing gas with average RMS velocity $\\sim60$\\,km\\,s$^{-1}$ and flow velocity $\\sim-10$\\,km\\,s$^{-1}$. Each phase has a different ionization (log $\\xi=-0.86, 0.87, 2.56$) and column density ($N_H=0.89\\times10^{20}, 2.41\\times10^{20}, 6.03\\times10^{20}$\\,cm$^{-2}$ respectively). Note that one continuous distribution of gas is unlikely, but not ruled out. The low flow velocity of the warm absorber phases can be partially explained by assuming the low ionization phases to be interstellar gas. The lower ionization phases are expected to contribute significantly to the UV spectrum, implying that the UV and X-ray absorbers are connected. \\newline The best fit to the RGS data requires five significant emission lines due to O\\,VIII Ly$\\alpha$ (18.9\\AA), O\\,VII(f) (22.1\\AA), N\\,VI(i) (29.0\\AA), N\\,VII Ly$\\alpha$ (24.7\\AA) and C\\,VI Ly$\\alpha$ (33.5\\AA). By examining the outflow velocities and density restrictions on these emission lines we have shown that they originate in two separate regions, the BLR and the NLR. It is possible that these regions and the absorber are connected. The outflowing broad line region expands as it travels out from the core, maintaining a similar ionization level. This gas then picks up dust from the torus, and slows down. As the gas continues to expand it then displays the characteristics of the NLR and the diffuse, dusty warm absorber. UV observations suggest a disturbed host galaxy. A tidal stream or dust and gas accumulating in the inner regions of the host galaxy offer an alternative origin for the low ionization absorber." }, "0808/0808.1289_arXiv.txt": { "abstract": "Observations using the {\\it Spitzer Space Telescope} provided the first detections of photons from extrasolar planets. {\\it Spitzer} observations are allowing us to infer the temperature structure, composition, and dynamics of exoplanet atmospheres. The {Spitzer} studies extend from many hot Jupiters, to the hot Neptune orbiting GJ\\,436. Here I review the current status of {\\it Spitzer} secondary eclipse observations, and summarize the results from the viewpoint of what is robust, what needs more work, and what the observations are telling us about the physical nature of exoplanet atmospheres. ", "introduction": "The powerful astrophysical leverage provided by transits enables us to study extrasolar planets directly, i.e., by detection of their emergent radiation. The {\\it Spitzer Space Telescope} has provided the bulk of these detections. The first {\\it Spitzer} measurements of exoplanet secondary eclipses were announced in 2005. Two independent groups (Charbonneau et al. 2005; Deming et al. 2005) measured eclipses for two different planets, using two different {\\it Spitzer} instruments, and obtained very similar results (Figure~1). \\begin{figure}[ht] \\begin{center} \\includegraphics[width=2.8in]{f1.eps} \\caption{First detections of exoplanet thermal emission using the {\\it Spitzer Space Telescope}. Plotted are the secondary eclipses of TrES-1 at 8\\,$\\mu$m (top, Charbonneau et al. 2005), and HD\\,209458b at 24\\,$\\mu$m (bottom, Deming et al. 2005)}. \\label{fig1} \\end{center} \\end{figure} Since each eclipse was independently measured to $\\sim 6\\sigma$ significance, exoplanet thermal emission was securely detected. The discovery of transits in HD\\,189733b (Bouchy et al. 2005) provided an opportunity to measure exoplanet thermal emission at higher signal-to-noise ratio. Initial observations of HD\\,189733b at 16\\,$\\mu$m (Deming et al. 2006) showed an eclipse of the planet at 32$\\sigma$ significance (Figure~2), and subsequent work using the IRAC instrument has detetcted the planet's flux to 60$\\sigma$ precision at 8\\,$\\mu$m (Knutson et al. 2007). This extraordinary level of precision in measuring exoplanet thermal emission allows many intersting studies that could hardly have been imagined when the first extrasolar planets were detected by radial velocity studies. \\begin{figure}[ht] \\begin{center} \\includegraphics[width=2.2in]{f2.eps} \\caption{Eclipse of HD\\,189733b at 16\\,$\\mu$m (Deming et al. 2006)}. \\label{fig2} \\end{center} \\end{figure} In this review I summarize highlights from {\\it Spitzer} secondary eclipse measurements, with some discussion of transmission spectroscopy during transit. The quality of work in this field has been uniformly high, but I will summarize the results from the viewpoint of what is robust, and what I believe needs more work and clarification. ", "conclusions": "" }, "0808/0808.0070_arXiv.txt": { "abstract": "{ The growth of supermassive black holes (SMBH) through accretion is accompanied by the release of enormous amounts of energy which can either be radiated away, as happens in quasars, advected into the black hole, or disposed of in kinetic form through powerful jets, as is observed, for example, in radio galaxies. Here, I will present new constraints on the evolution of the SMBH mass function and Eddington ratio distribution, obtained from a study of AGN luminosity functions aimed at accounting for both radiative and kinetic energy output of AGN in a systematic way. First, I discuss how a refined Soltan argument leads to joint constraints on the mass-weighted average spin of SMBH and of the total mass density of high redshift ($z\\sim 5$) and ``wandering'' black holes. Then, I will show how to describe the ``downsizing'' trend observed in the AGN population in terms of cosmological evolution of physical quantities (black hole mass, accretion rate, radiative and kinetic energy output). Finally, the redshift evolution of the AGN kinetic feedback will be briefly discussed and compared with the radiative output of the evolving SMBH population, thus providing a robust physical framework for phenomenological models of AGN feedback within structure formation. ", "introduction": "Black holes in the local universe come into two main families according to their size, as recognized by the strongly bi-modal distribution of the local black hole mass function (see Fig.~\\ref{fig:mf}). While the height, width and exact mass scale of the stellar mass peak should be understood as a by-product of stellar (and binary) evolution, and of the physical processes that make supernovae and gamma-ray bursts explode, the supermassive black holes one is the outcome of the cosmological growth of structures and of the evolution of accretion in the nuclei of galaxies, likely modulated by the mergers these nuclear black holes will experience as a result of the hierarchical galaxy-galaxy coalescences. \\begin{figure*}[t!] \\resizebox{\\hsize}{!}{\\includegraphics[clip=true]{shankar_mf.ps}} \\caption{ \\footnotesize The local black hole mass function, plotted as $M \\times \\phi_M$, in order to highlight the location and height of the two main peaks. The stellar mass black holes peak has been drawn assuming a log-normal distribution with mean mass equal to 5 solar masses, width of 0.1 dex and a normalization yielding a density of about $1.1 \\times 10^{7}$ $M_{\\odot}$ Mpc$^{-3}$ \\citep{fukugita:04}. The supermassive black hole peak, instead, contribute to an overall density of about $4.3 \\times 10^{5}$ $M_{\\odot}$ Mpc$^{-3}$ \\citep{merloni:08}} \\label{fig:mf} \\end{figure*} In the recent literature, it has become customary to introduce the works on cosmological aspects of AGN astrophysics by referring to the strong role they most likely play in the galaxy formation process throughout cosmic history. Indeed, a new paradigm has emerged, according to which the feedback energy released by growing supermassive black holes (i.e. AGN) limits the stellar mass growth of their host galaxies in a fundamental, generic, but yet not fully understood fashion. The strongest observational evidence for such a schematic picture emerged in the last decade. The search for the local QSO relics via the study of their dynamical influence on the surrounding stars and gas carried out since the launch of the {\\it Hubble Space Telescope} \\citep[see e.g.][and references therein]{richstone:98,ferrarese:08} led ultimately to the discovery of tight scaling relations between SMBH masses and properties of the host galaxies' bulges \\citep{gebhardt:00,ferrarese:00,tremaine:02,marconi:03}, clearly pointing to an early co-eval stage of SMBH and galaxy growth. A second piece of evidence comes from X-ray observations of galaxy clusters, showing that black holes are able to deposit large amounts of energy into their environment in response to radiative losses of the cluster gas. From studies of the cavities, bubbles and weak shocks generated by the radio emitting jets in the intra-cluster medium (ICM) it appears that AGN are energetically able to balance radiative losses from the ICM in the majority of cases \\citep[see][and references therein]{birzan:08}. Nevertheless, the physics of AGN heating in galaxy cluster is still not well established, neither have the local scaling relations proved themselves capable to uniquely determine the physical nature of the SMBH-galaxy coupling. As a consequence, a large number of feedback models have so far been proposed which can reasonably well reproduce these relations. From the observational point of view, the crucial test for most models will be a direct comparison with the high-redshift evolution of the scaling relations. There is, however, another benchmark, based on existing data, all models have to be tested upon: the evolution of the SMBH mass function and of the predicted energy output (either in radiative or kinetic form) needed to offset gas cooling and star formation in galaxies. Here I present our recent attempt to reconstruct the history of SMBH accretion in order to follow closely the evolution of the black hole mass function, as needed in order to test various models for SMBH cosmological growth as well as those for the black hole-galaxy co-evolution. Similar to the case of X-ray background synthesis models, where accurate determinations of the XRB intensity and spectral shape, coupled with the resolution of this radiation into individual sources, allow very sensitive tests of how the AGN luminosity and obscuration evolve with redshift, we have argued that accurate determinations of the local SMBH mass density and of the AGN (bolometric) luminosity functions, coupled with accretion models that specify how the observed AGN radiation translates into a black hole growth rate, allow sensitive tests of how the SMBH population (its mass function) evolves with redshift. By analogy, we have named this exercises `AGN synthesis modelling' \\citep{merloni:08}. In performing it, we have taken advantage of the fact that the cosmological evolution of SMBH is markedly simpler than that of their host galaxies, as individual black hole masses can only grow with time, and SMBH do not transform into something else as they grow. Moreover, by identifying active AGN phases with phases of black holes growth, we can follow the evolution of the population by solving a simple continuity equation, where the mass function of SMBH at any given time can be used to predict that at any other time, provided the distribution of accretion rates as a function of black hole mass is known (see \\S~\\ref{sec:integral}). \\begin{figure*}[t!] \\begin{tabular}{cc} \\resizebox{0.48\\hsize}{!}{\\includegraphics[clip=true]{rhobh_mass_083.ps}}& \\resizebox{0.48\\hsize}{!}{\\includegraphics[clip=true]{zev_lambda_083_masses.ps}}\\\\ \\end{tabular} \\caption{\\footnotesize Redshift evolution of the SMBH mass density (left) and average Eddington ratio (right), calculated for different BH mass bins (in solar mass units). In the right hand plot, black lines and grey shaded area represent the overall (mass-wighted) average Eddington ratio.} \\label{fig:zev_mdot} \\end{figure*} In order to carry out our calculation, we assumed that black holes accrete in just three distinct physical states, or ``modes'': a radiatively inefficient, kinetically dominated mode at low Eddington ratios (LK, the so-called ``radio mode'' of the recent literature), and two modes at high Eddington ratios: a purely radiative one (radio quiet, HR), and a kinetic (radio loud, HK) mode, with the former outnumbering the latter by about a factor of 10. Such a classification is based on our current knowledge of state transitions in stellar mass black hole X-ray binaries as well as on a substantial body of works on scaling relations in nearby SMBH. It allows a relatively simple mapping of the observed luminosities (radio cores, X-ray and/or bolometric) into the physical quantities related to any growing black hole: its accretion rate and the released kinetic power. In this work I will focus on just a few specific aspects of the derived SMBH evolution, in particular on the redshift evolution of the mass function and Eddington ratio distribution, on the constrains we put on the mass-weighted average spin of the SMBH population, and on the kinetic energy output of growing black holes. A more detailed discussion of the methodology, as well as a wider exploration of our results can be found in \\citet{merloni:08}. All results will be shown accounting for the intrinsic uncertainties of the adopted luminosity functions. We estimated that these uncertainties can be evaluated by comparing different analytic parametrization of the same data sets; specifically, we adopted the LDDE and MPLE parametrization for the hard X-ray luminosity function of \\citet{silverman:08}, and two alternative parametrizations for the flat-spectrum radio luminosity function of \\citet{dunlop:90} and \\citet{dezotti:05}. ", "conclusions": "I have outlined some recent results of our work aimed at pinning down as accurately as possible the cosmological evolution of active galactic nuclei and of the associated growth of the supermassive black holes population. In particular, I have focused here on the global (integrated) constraints on the mass-weighted average spin of SMBH, and I have discussed in some details the specific ways in which these are tighten to the very interesting open issues regarding the population of high-redshift SMBH and that of black holes ejected from galactic nuclei due to gravitational wave recoil in merger events. I have also discussed a few generic properties of the kinetic energy output of growing black holes, emphasizing the importance of a late phase of low-luminosity, jet-dominated accretion onto the most massive objects. The richness of details we have been able to unveil demonstrates that times are ripe for comprehensive unified approaches to the multi-wavelength AGN phenomenology. At the same time, our results should serve as a stimulus for semi-analytic and numerical modelers of structure formation in the Universe to consider more detailed physical models for the evolution of the black hole population." }, "0808/0808.0300_arXiv.txt": { "abstract": "In this paper we introduce a new linear filtering technique, the so-called matrix filters, that maximizes the signal-to-interference ratio of compact sources of unknown intensity embedded in a set of images by taking into account the cross-correlations between the different channels. By construction, the new filtering technique outperforms (or at least equals) the standard matched filter applied on individual images. An immediate application is the detection of extragalactic point sources in Cosmic Microwave Background images obtained at different wavelengths. We test the new technique in two simulated cases: a simple two-channel case with ideal correlated color noise and more realistic simulations of the sky as it will be observed by the LFI instrument of the upcoming ESA's Planck mission. In both cases we observe an improvement with respect to the standard matched filter in terms of signal-to-noise interference, number of detections and number of false alarms. ", "introduction": "\\IEEEPARstart{T}{he} detection of faint pointlike sources is a task that is common to many branches of Astronomy, from the search for protostars in gas-rich nebulae to the study of active galactic nuclei in the confines of the observable universe. Since the angular size of these objects is smaller than the angular resolution of the telescopes that are used to observe them, they appear as \\emph{point sources} with the shape of the telescope point spread function. A case of particular interest is the detection of extragalactic point sources (EPS) in microwave wavelengths. In the microwave range of the electromagnetic spectrum, the sky is dominated by the so-called Cosmic Microwave Background (CMB) radiation, a relic of the hot and dense first moments of the universe. The study of the CMB is one of the hottest research topics in modern Cosmology. For a short review on the CMB, see~\\cite{cmbrev06}. The CMB signal is mixed with other signals of astrophysical origin, mainly the emission from our own Galaxy and from a large number of extragalactic objects including radio galaxies, dusty galaxies and galaxy clusters. For the typical angular resolution of current CMB experiments, ranging from a few arcmin to one degree, most of these extragalactic objects appear as point sources. From the standpoint of CMB, these point sources are contaminants that must be removed; from the standpoint of extragalactic Astronomy, however, they provide a valuable source of information, particularly at microwave wavelengths where the properties of sources are poorly understood~\\cite{bluebook}. In both cases, techniques for the detection of faint point sources are needed. For single images taken at a given wavelength the problem is equivalent to the general problem of detecting a number of objects, all of them with a known waveform but unknown positions and intensities, embedded in additive noise (not necessarily white). In the field of microwave Astronomy, wavelet techniques~\\cite{vielva01,vielva03,MHW206,wsphere,NEWPS07}, Bayesian approaches~\\cite{hobson03,powell08}, matched filters~\\cite{tegmark98,barr03,can06} and other related linear filtering techniques~\\cite{sanz01,naselsky02,herr02b,herr02c,can04a,can05a,can05b} have proved to be useful. The common feature of all these techniques is that they rely on the prior knowledge that the sources have a distinctive spatial behaviour (i.e.~a known spatial profile, plus the fact that they appear as compact objects as opposed to `diffuse' random fields) that helps to distinguish them from the noise. Most of the current CMB experiments are able to observe the sky at several different wavelength bands. In particular, the Wilkinson Microwave Anisotropy Probe (WMAP)~\\cite{wmap0,hinshaw07} is observing the sky at 23, 33, 41, 61 and 94 GHz and the upcoming ESA's Planck satellite~\\cite{planck_tauber05} will observe the sky in nine frequency channels ranging from 30 to 857 GHz. Multi-wavelength data can be used to improve the detection/separation of astrophysical components. For example, CMB can be separated from Galactic dust and synchrotron emission using well-established methods; a non-exhaustive list of them include Independent Component Analysis~\\cite{fastica02,fastica07}, Maximum Entropy Method~\\cite{MEM97,vlad02,barreiro_MEM04}, Internal Linear Combination~\\cite{wmap0,eriksen04} and Wiener filtering~\\cite{bouchet99}, among others. Moreover, specific techniques for the detection of compact sources whose spectral behaviour is well known have been proposed in the literature; a typical example is the detection of the so-called Sunyaev-Zel'dovich effect due to galaxy clusters~\\cite{herr02a,chema02,schaf06,melin06}. All the previous component separation techniques, however, are in trouble when dealing with extragalactic point sources. EPS form a very heterogeneous population constituted by a large number of objects with very different physical properties, from radio-emitting active galactic nuclei to dusty star-forming galaxies. Since each source is a unique object, with an spectral emission law that is different from the spectral emission law of any other, the component separation problem is underdetermined: the number $N$ of components to be separated is much larger than the number $M$ of channels available. Attempts to simplify the problem by grouping the extragalactic sources into classes of objects with similar spectral behaviour are in most cases unsatisfactory. So far we have considered two approaches: on the one hand, it is still possible to work channel by channel, separately, by using the filtering techniques above mentioned. But in that case a valuable fraction of the information that multi-wavelength experiments can offer is wasted. On the other hand, standard multi-wavelength component separation techniques are impracticable because the number of physical components involved is too high. An intermediate approach is to design filters that are able to find compact sources thanks to their distinctive spatial behaviour while at the same time do incorporate some multi-wavelength information, without pretending to achieve a full component separation. In this paper we propose a new filtering technique that makes no assumptions about the spectral behaviour of the sources, but that makes use of some multi-wavelength considerations, namely: \\begin{itemize} \\item When a source is found in one channel, it will be also present in the same position in all the other channels. \\item The spatial profile of the sources may differ from channel to channel, but it is a priori known. For example, for a microwave experiment the source profile is equal to the antenna response of the experiment's radiometers\\footnote{In general, of its detectors. Some microwave experiments use bolometers instead of radiometers, but the distinction is irrelevant for the purposes our discussion.}, that is well known. \\item The second order statistics of the background in which the sources are embedded is well known or it can be directly estimated from the data by assuming that point sources are sparse. \\end{itemize} \\noindent We find that the new technique we propose takes the form of a matrix of filters that can be applied to any number $M$ of channels. In the case of a single channel, the matrix of filters defaults to the standard matched filter. In section~\\ref{sec:math} we will introduce the new formalism. In section~\\ref{sec:tests} we will illustrate the new method with two different tests: a simple toy model and a more realistic simulation that corresponds to a small region of the sky as it will be observed by the upcoming Planck mission. Finally, in section~\\ref{sec:conclusions} we will draw some conclusions. ", "conclusions": "\\label{sec:conclusions} In this work we have introduced a new filtering technique, the matrix filters, that can be of utility for the detection of extragalactic point sources in multi-wavelength experiments of the Cosmic Microwave Background. Matrix filters are designed to maximize the signal-to-interference ratio of compact sources embedded in a set of images (``channels'') by taking into account the cross-correlations between the different channels. For the case of a single-channel experiment and circularly symmetric sources, the matrix of filters defaults to the standard matched filter. For the case of $M$ totally uncorrelated channels and circularly symmetric sources, the matrix of filters becomes a diagonal matrix whose non-zero elements are the matched filters corresponding to each channel. In a general case, the non diagonal elements of the matrix are non zero. Since the matrix filters contain as a particular case the matched filter and they are designed to maximize the signal-to-interference ratio, then the individual amplifications for each channel (defining by amplification the quotient between the signal-to-interference ratios after and before filtering) are greater or (in the worst case) equal to the amplifications obtained by the standard matched filters. We have tested the matrix filters in two simulated cases: a simplistic two-channel case with ideal noise and a more realistic simulation of the sky as it will be observed by the LFI instrument of the upcoming ESA's Planck mission. In the first test we observe that matrix filters clearly outperform the standard matched filter for one of the channels, while for the other channel their performance is very similar to that of the matched filter. Other simulation tests with different parameters we have carried out show a similar behaviour: there is a significant gain for at least one of the channels with respect to standard matched filters. For the second test case we have considered one single patch of the sky, observed at 30, 44 and 70 GHz. The purpose of this test was just to give an example of how the matrix filters work in a more realistic case. In this example, matrix filters outperform the matched filter both in the flux limit they can reach and in the ratio between false alarms and true detections. This result seems to indicate that matrix filters could help to obtain better catalogues of extragalactic point sources in future CMB experiments. However, this result has been obtained for a single simulation of a small portion of the sky, and therefore it may not be extrapolable to the whole sky or to any other experiment. The study of the application of matrix filters to the whole microwave sky for the Planck mission is the subject of a future work." }, "0808/0808.2135.txt": { "abstract": "We present a time-dependent multi-zone code for simulating the variability of Synchrotron-Self Compton (SSC) sources. The code adopts a multi-zone pipe geometry for the emission region, appropriate for simulating emission from a standing or propagating shock in a collimated jet. Variations in the injection of relativistic electrons in the inlet propagate along the length of the pipe cooling radiatively. Our code for the first time takes into account the non-local, time-retarded nature of synchrotron self-Compton (SSC) losses that are thought to be dominant in TeV blazars. The observed synchrotron and SSC emission is followed self-consistently taking into account light travel time delays. At any given time, the emitting portion of the pipe depends on the frequency and the nature of the variation followed. Our simulation employs only one additional physical parameter relative to one-zone models, that of the pipe length and is computationally very efficient, using simplified expressions for the SSC processes. The code will be useful for observers modeling GLAST, TeV, and X-ray observations of SSC blazars. ", "introduction": "} In blazars, radio loud active galaxies with their relativistic jets pointing close to our line of sight \\citep{blandford78}, the observed radiation is dominated by relativistically beamed emission from the sub-pc base of the jet. The blazar spectral energy distribution (SED) consists of two components. The first one, peaking at sub-mm to X-ray energies is almost certainly due to synchrotron radiation, while the second one peaking at MeV to TeV energies is believed to be of inverse Compton (IC) nature, with both components produced by the same population of relativistic electrons. The nature of the IC-scattered seed photons is still not clear, with both external optical-UV photons from the broad line region \\citep{sikora94} and IR photons from the putative molecular torus \\citep{blazejowski00}, as well as synchrotron photons (SSC, e.g. Maraschi, Ghisellini \\& Celotti 1992) contributing. It is believed that in the case of powerful blazars peaking at MeV to GeV energies, external seed photons from the broad line region dominate the IC scattering, while for weaker lineless blazars, peaking at $\\sim$ TeV energies, SSC is the dominant emission mechanism. Recent observational results (e.g. D'Arcangelo et al. 2007; Marscher et al. 2008), however, place the blazar emission site beyond the broad line region, lending support to the possibility that even in powerful blazars the GeV emission process may be pure SSC. For a review of leptonic models, as well as hadronic models for blazar emission (e.g. Aharonian 2000) see \\citet{boettcher06}. Due to the small angular size of the blazar emission region, it is not possible to spatially resolve the emitting region. Because of this, information about the structure of the emitting source can be obtained only through multiwavelength variability studies. Particularly telling is the variability of the emission produced by the highest energy electrons, because these electrons lose energy very quickly and exist only close to the sites where they have been produced. The goal of multiwavelength variability campaigns, involving in many cases observations from radio up to TeV energies, is to study the characteristics of blazar variability, such as correlations and/or time delays between different energies, spectral characteristics of the observed variability, and the amplitude of variability as a function of energy. Most notable amongst blazars are the so called TeV blazars for which the synchrotron emission peaks at X-ray energies and the SSC emission peaks at TeV energies, as they present the active galaxies producing the highest confirmed electron energies. Variations of TeV blazars in these two bands can be extremely rapid (TeV doubling times as short as a few min; Aharonian et al. 2007), suggesting highly relativistic sub-pc scale flows (Doppler factors $\\delta\\sim 50$; e.g. Begelman, Fabian \\& Rees 2008) that decelerate substantially \\citep{georganopoulos03,ghisellini05} to match the much slower speeds required by VLBI observations \\citep{piner04, piner08}. The TeV and X-ray variations are usually well correlated (e.g. Fossati et al. 2008; Maraschi et al. 1999; Sambruna et al. 2000), as expected, because they present variations by the same electron population. Usually, the lower energy emission within each of these bands peaks with small time delays relative to the higher energy emission (e.g. Fossati et al. 2000), while the X-ray and TeV spectra become harder with increasing flux (e.g. Takahashi et al. 1996). In certain cases, however, the X-ray and TeV variability do not seem to be correlated in a simple way (e.g. Aharonian et al. 2005). An intriguing variability pattern is that of the so-called `orphan' flares, rare TeV flares that are not accompanied by X-ray flares (e.g. Krawczynski et al. 2004, B\\l a\\.zejowski et. al. 2005). While the correlated X-ray - TeV flares can be understood through an increase of the high energy emitting electrons, orphan TeV flares defy such a straightforward explanation. Models of blazar emission to date have, for the most part, been in some form of homogeneous one zone models (e.g. Mastichiadis \\& Kirk 1997; Krawczynski, Coppi, \\& Aharonian 2002). Such models, although appropriate for modeling the steady-state emission of a source, cannot simulate variability faster than the zone light crossing time. The basic limitation of one zone models stems from the fact that the high energy variability of both the synchrotron and SSC components is produced by high energy electrons with cooling times shorter than the light crossing time. Even if we assume that a disturbance in the radiating plasma (e.g. a higher density) instantaneously propagates across the zone, the received radiation would be smeared out for timescales shorter than the light crossing time, due to light travel time delays from different parts of the source (\\S \\ref{section:testing}; also Chiaberge \\& Ghisellini 1999), and no variability faster than the light crossing time would be observed. One, therefore, cannot use one zone models to infer the source structure from high energy variability. Inhomogeneous variability models of increasing degree of sophistication have attempted to overcome the problems of one-zone models. The basic idea is to overcome the unphysical instantaneous injection throughout the source by adopting a specific geometry for the plasma flow that includes an inlet for injecting the radiating plasma. Variations in the injected plasma propagate and produce variations in the emissivity. Calculations of the received emission that take into account the light-travel times that radiation from different parts of the source takes to reach the observer, produce light curves that, at least, do not violate causality. How physically realistic these light curves are depends on the approximations used and on the characteristics of the source to be modeled. For example, synchrotron and IC losses from photons external to the source are local processes in the sense that, at a given point in the flow, the energy loss rate only depends on the local magnetic field and external photon field energy density, and not on the photon production throughout the source. This is not the case with SSC losses, because synchrotron photons produced throughout the source at earlier times - to take into account the light travel time from one point of the source to another - contribute to the photon energy density responsible for the SSC losses and to the emissivity at a given point and time in the source. To properly model sources like TeV blazars, in which SSC losses are important or even dominant, these considerations have to be taken into account. There have been a few attempts during the last fifteen years to take these spatial considerations into account. \\cite{gomez94} considered a conical jet with a constant bulk Lorentz factor flow in which the electron plasma and the magnetic field undergo adiabatic evolution only and calculated the radio variability induced by a shock wave propagating along the jet. \\cite{georganopoulos98a,georganopoulos98b} studied a parabolic jet that hydrodynamically accelerates and focuses to a conical geometry, and by following the synchrotron energy losses of the emitting electrons reproduced the radio to X-ray light curves of the X-ray bright blazar PKS 2155-304. This resulted in a frequency-dependent source size, in agreement with the fact that the variability timescale of synchrotron radiation increases with decreasing frequency. It also reproduced the usually observed soft lags (variations at soft X-rays being preceded by variations at hard X-rays) and the counterclockwise X-ray flux - X-ray spectral index loops (e.g. Takahashi 1996, Maraschi 1999, Kataoka 2000, Ravasio 2004), both manifestations of radiative cooling dominating the energetics of the high energy electrons. \\cite{kirk98} developed a semi-analytical model in which low energy electrons are injected in a zone where they undergo acceleration and eventually escape. The acceleration zone is assumed to move with a certain velocity, leaving behind the freshly accelerated electrons that cool through synchrotron radiation. Variations in the injection rate of low energy electrons in the acceleration zone result in variations of the emissivity, which are integrated over the volume of the source, taking into account time delays, to produce the observed multifrequency synchrotron light curves. This model includes a treatment of particle acceleration and it is able to reproduce the uncommon hard lags (variations at hard X-rays preceded by variations at soft X-rays) and clockwise X-ray flux - X-ray spectral index loops \\citep{zhang02,ravasio04}, both manifestations of electrons still accelerating, just before reaching the maximum electron Lorentz factor, where the acceleration and radiative loss timescales are comparable. It does not include, however, SSC considerations. \\cite{chiaberge99} studied the synchrotron and SSC emission, from a homogeneous one zone model in which they assumed an instantaneous plasma injection, but taking into account the time delays with which the external observer would observe the variability (a similar approach was also taken by Kataoka et al. 2000). They also studied a case similar to that of Kirk et al. (1998), but without treating particle acceleration, by splitting the source into smaller one zone models that evolved autonomously, in the sense that ({\\sl i}) the SSC emission inside any one of their single zones uses as seed photons only the synchrotron photons produced in that zone and ({\\sl ii}) the SSC energy losses in every zone are caused only by the synchrotron photons produced in that zone. This simplified approach is a good approximation for following the energetics of the electrons if the source is synchrotron dominated (because the SSC losses, although inappropriately calculated, are negligible), but does not produce realistic SSC light curves, because it does not calculate the emission due to upscattering synchrotron photons produced in other parts of the source in retarded times. A significant improvement was introduced by \\cite{sokolov04} who incorporated in the calculation of the SSC emission from a given location in an inhomogeneous source the synchrotron photons produced throughout the source in retarded times. This produces accurate SSC light curves, provided that the SSC losses that were still treated as a local process are negligible. In a follow up paper, \\cite{sokolov05}, considered also external Compton photons from the broad line region and the molecular torus. The challenge for inhomogeneous multi zone models for sources such as the TeV emitting blazars is the calculation of the non-local, time-retarded SSC losses induced by photons produced in other parts of the source. Here, we present such an inhomogeneous model that, for the first time, takes into account the non-local, time-delayed source emission on the SSC losses. We assume that a power law of relativistic electrons is injected at the inlet of a pipe, and that the electrons flow downstream and cool radiatively. Variations in the injected electron distribution propagate downstream and manifest themselves as frequency dependent variability. This allows us to model high energy multiwavelength variability in a self-consistent manner. In \\S \\ref{section:onezone} we describe the one zone model, which we use as a building block for the multizone model, and we show that, by construction, one zone models cannot simulate variability produced by high energy electrons with radiative cooling time shorter than the electron light crossing time from the single zone. In \\S \\ref{section:multizone} we describe our multizone, pipe-geometry model with emphasis on the coupling between subsequent zones and on the calculation of the local photon field due to non-local, time-delayed emission throughout the source. This is followed in \\S \\ref{section:results} by a comparison of the code with analytical results and a series of case studies. We conclude in \\S \\ref{section:discussion} with a discussion of additional considerations that can be used as starting points for future work. ", "conclusions": "} We presented a multizone code that for the first time takes into account the non-local, time-retarded nature of SSC losses. This code is currently the only multizone model that incorporates the non-local, time-delayed SSC losses, and as such is uniquely suitable for modeling the results of multiwavelength campaigns at radio, optical, X-ray and $\\gamma$-ray energies, with the additional constraints in the critical and unexplored for TeV blazars GeV GLAST regime. As we argued, the results of one zone codes for the critical high energy regime of both the synchrotron and SSC components are problematic, and should not be used to infer the physical conditions in the source through variability modeling. We described our multizone code, tested it successfully against known analytical results, and presented a small number of variability case studies. The case studies we presented, although based on the same underlying steady-state configuration, exhibited very different variability patterns. This means that detailed modeling of broadband SEDs and simultaneous multiwavelength variability can be used to infer what is actually the cause of a given observed variability pattern, providing reliable constraints on the particle acceleration taking place. Orphan flares can be reproduced assuming an increase of the injection of the low energy electrons, but not assuming the injection of a very high energy electron population, as we also showed analytically. The fact that this plausible variation cannot produce orphan flares significantly narrows the parameter space for events that can produce such events, possibly in agreement with their observed scarcity. The code we described can run with a typical workstation in a reasonable time of at most a few minutes at a resolution of $\\sim 10$ bins per decade of observing frequency, $\\sim 10$ bins per decade of electron energy, and $\\sim 50$ zones. To achieve this we employed a pipe geometry, and adopted an energy conserving $\\delta$-function approximation for the SSC emissivity, as well as a step function approach to take into account the change from the Thomson to Klein-Nishina IC scattering cross-sections. Adopting these approximations is problematic for situations where IC scattering of narrow photon distributions (e.g. line emission from the broad line region or even a blackbody spectrum characterized by a typical photon energy $\\epsilon_0$) is important. In this case the adoption of the step function cross section description would create a strong artificial feature on the EED localized at the transition from the Thomson to the Klein-Nishina regimes at $\\gamma\\propto 1/\\epsilon_0$, which would then propagate to the emitted spectra through the $\\delta$-function IC emissivity. For SSC systems, however, where the seed photons are spread over many decades in energy, the resulting spectra are good approximations of those produced using the full expressions for the synchrotron and SSC emissivities as well as the full Klein-Nishina cross section. Including the above considerations, as well as the processes of synchrotron opacity and pair production through $\\gamma$-ray absorption within the source, would increase the execution time up to levels marginally comfortable for the typical workstation. Most probably, such an extension of the code would require parallelization. A more desirable upgrade of the code would drop the assumption of no lateral gradients in the plasma characteristics by switching to a two-dimensional geometry, in which the electron distribution and the SSC photon energy density are allowed to change laterally to the flow direction. Such considerations may be relevant to the recently observed $\\sim 0.75 $ days delay between the IR and the X-ray variability in 3C 273 \\citep{mchardy07}. These authors argued that the delay may be attributed to the time it takes for the SSC photon energy density to built up as the SSC photons are transversing the cross section of the flow. This upgrade will scale the computation time roughly by $N^2$, increasing it from $\\sim$ few minutes to $\\sim$ several hours. We note here that our formalism can be extended to treat velocity profiles in term of the decelerating flow \\citep{georganopoulos03} or the spine sheath model \\citep{ghisellini05} that have developed to address the lack of superluminal motions in TeV blazars (e.g. Piner \\& Edwards 2004). Another upgrade that can be incorporated in the existing code, this time with a minimal computational overhead, is that of a zone for particle acceleration, following the formalism of \\cite{kirk98}. In this case, in the first zone of the model, low energy electrons will be injected and allowed to accelerate while suffering radiative losses due to synchrotron and non-local SSC. These particles will subsequently escape into the pipe and flow downstream. This configuration will require a different numerical scheme for the acceleration zone, since there most particles are advected upward in energy space, but there is a possibility, in a time-dependent scenario, of the highest energy particles being advected downward, while the rest of the electrons are still advected upwards. The benefit of including particle acceleration in the code is that it will allow us to study cases of hard lags/ counterclockwise loops in the X-ray hardness - X-ray flux diagrams thought to result when acceleration and loss timescales are comparable. Such a code could be used to model the observed curved X-ray spectra of high peak frequency blazars in the framework of episodic particle acceleration \\citep{perlman05}." }, "0808/0808.2582_arXiv.txt": { "abstract": "The muon charge ratio of ultrahigh energy cosmic rays may provide information to detect the composition of the primary cosmic rays. We propose to extract the charge information of high energy muons in very inclined extensive air showers by analyzing their relative lateral positions in the shower transverse plane. ", "introduction": " ", "conclusions": "" }, "0808/0808.2061_arXiv.txt": { "abstract": "Firmani et al. proposed a new Gamma Ray Burst (GRB) luminosity relation that showed a significant improvement over the $L_{iso} - E_{peak}$ relation. ($L_{iso}$ is the isotropic peak luminosity and $E_{peak}$ is the photon energy of the spectral peak for the burst.) The new proposed relation simply modifies the $E_{peak}$ value by multiplying it by a power of $T_{0.45}$, where $T_{0.45}$ is a particular measure of the GRB duration. We begin by reproducing the results of Firmani for his 19 bursts. We then test the Firmani relation for the same 19 bursts except that we use independently measured values for $L_{iso}$, $T_{0.45}$, and $E_{peak}$, and we find that the relation deteriorates substantially. We further test the relation by using 60 GRBs with measured spectroscopic redshifts, and find a relation that has a comparable scatter as the original $L_{iso} - E_{peak}$ relation. That is, a much larger sample of bursts does not reproduce the small scatter as reported by Firmani et al. Finally, we investigate whether the Firmani relation is improved by the use of any of 32 measures of duration (e.g., $T_{90}$, $T_{50}$, $T_{90}/N_{peak}$, the fluence divided by the peak flux, $T_{0.30}$, and $T_{0.60}$) in place of $T_{0.45}$. The quality of each alternative duration measure is evaluated with the root mean square of the scatter between the observed and fitted logarithmic $L_{iso}$ values. Although we find some durations yield slightly better results than $T_{0.45}$, the differences between the duration measures are minimal. We find that the addition of a duration does not add any significant improvement to the $L_{iso} - E_{peak}$ relation. We also present a simple and direct derivation of the Firmani relation from {\\it both} the $L_{iso} - E_{peak}$ and Amati relations. In all we conclude that the Firmani relation neither has an independent existence nor does it provide any significant improvement on previously known relations that are simpler. ", "introduction": "To date, eight separate luminosity relations have been identified for long duration Gamma-Ray Bursts (GRBs) (Schaefer and Collazzi, 2007). These relations correlate a burst's peak bolometric luminosity with various light curve and spectral parameters. The possible utility of these relations is that we can then use GRBs as tracers of the high-redshift universe. One of these eight relations is the $L_{iso} - E_{peak}T_{0.45}$ relation proposed by Firmani et al. (2006, hereafter named the Firmani relation). Here, $E_{peak}$ describes the peak of the $E*F(E)$ curve (proportional to $\\nu F_{\\nu}$ ), which is the photon energy of the peak spectral flux; $L_{iso}$ is the isotropic luminosity of the burst measured bolometrically (1-10,000 kev in the burst rest frame). $T_{0.45}$ is the Reichart definition of a GRB time duration (Reichart et al. 2001) where the duration is the total time interval of the brightest bins in the light curve that contains 45$\\%$ of the burst fluence. The Firmani relation was presented as an improvement over the $L_{iso} - E_{peak}$ relation. Nineteen GRBs were used to demonstrate a tight correlation with a reduced chi-square of 0.7 over 16 degrees of freedom; the resulting luminosity relation being $L_{iso} \\propto {E_{peak}}^{1.62} {T_{0.45}}^{-0.49}$. The reported scatter in the Firmani relation is substantially smaller than those of most other GRB luminosity relations. This result offers the hope of substantial improvement in the accuracy of GRBs for cosmological distance measures. We have no physical reason to expect that the addition of a duration should make for a tighter relation. Nonetheless, we should look to see if we can get tighter luminosity relations from duration definitions other than $T_{0.45}$, as $T_{0.45}$ may not be the optimal duration to use. We have no physical reason to expect that any one definition of duration is best, while the particular choice of the Reichart definition was used only for historical reasons no longer of any relevance. For example, we could still use the Reichart definition, but measure the duration over a different percentage of the burst fluence. So perhaps the use of a duration based on 30\\% or 60\\% ($T_{0.30}$ or $T_{0.60}$) might be better. Alternative definitions of duration can be considered instead of the Reichart formulation. For example, we can try to define the duration as equalling the total fluence divided by the peak flux to get a sort of equivalent width; alternatively, we could use the familiar $T_{90}$ or $T_{50}$ durations. A wide variety of alternative durations can be defined, and we do not know which one will be optimal. In this Section 2, we first reproduce the Firmani relation for the original data of Firmani et al. (2006) as a test that we are using identical fitting procedures. Then we test the Firmani relation with a set of independent data for the same 19 bursts. A further test of the Firmani relation is made with a much larger sample of 60 bursts. In Section 3, we test many duration definitions in a Firmani-like relation to see which produces the `tightest' correlation. Section 4 contains a simple and forced derivation of the Firmani relation from two other luminosity relations. ", "conclusions": "In a recent independent preprint, Rossi et al. (2008) also examined the Firmani relation, in particular with a comparison to the Amati relation. They use an extended sample of 40 {\\it BeppoSAX} and {\\it Swift} bursts, with little overlap with our sample of 60 GRBs. Their best fit is somewhat different from those in Eqs 1, 4, or 7; with their fitted Firmani relation scaling as $L_{iso} \\propto E_{peak}^2 T_{0.45}^{-1}$. They realized that this Firmani relation is essentially identical to the Amati relation (Amati et al. 2002), which gives the isotropic energy emitted in gamma radiation over the whole burst duration as $E_{\\gamma ,iso} \\propto E_{peak}^2$. With the reasonable approximation that the total energy in the light curve equals the peak luminosity times the duration ($E_{\\gamma ,iso} \\approx L_{iso} T_{0.45}$), the Amati relation ($E_{\\gamma ,iso} \\propto E_{peak}^2$) is transformed into their Firmani relation ($L_{iso} \\propto E_{peak}^2 T_{0.45}^{-1}$). While our exponents in the Firmani relation are somewhat different, the Rossi derivation demonstrates that the Fermani relation has a physical basis that is close to that of the Amati relation. Rossi et al. (2008) further go on to show that the scatter in their Firmani relation is comparable to that in the Amati relation, which is another way of saying that the two relations are not independent. Here, we will directly derive the Firmani relation from {\\it both} the $L_{iso} - E_{peak}$ and Amati relations. Let us start with the relation $L_{iso} \\propto E_{peak}^{1.68}$ as given in Schaefer (2007). This can be rearranged as $L_{iso} \\propto E_{peak}^{1.9} (E_{peak}^2/L_{iso})^{-0.69}$. The Amati relation ($E_{\\gamma ,iso} \\propto E_{peak}^2$) can be inserted to get $L_{iso} \\propto E_{peak}^{1.9} (E_{\\gamma ,iso}/L_{iso})^{-0.69}$. Now we can select one of our duration definitions with $\\tau = S_{bolo}/P_{bolo}$. The ratio of fluence to peak flux will equal to the ratio of the burst energy and the peak luminosity, so we have $\\tau = E_{\\gamma ,iso}/L_{iso}$. This can now be substituted to obtain $L_{iso} \\propto E_{peak}^{1.9} \\tau^{-0.69}$. We see that we have just derived Eq. 8 with $\\xi = 1.9$ and $\\eta = -0.69$, values which are characteristic of the fitted Firmani relation (cf. Eq. 7). With this, we see that the Firmani relation has no independent existence because it is only a combination of two simpler relations. Thus, given any two of these three relations, we can derive the third. A question is which of these is more fundamental. The inherent problem with making this assessment is that it comes down to how one identifies the more fundamental of relations that really address different physics. By Occam's Razor, the more fundamental relation would be one that has the fewest parameters, while accurately and efficiently yielding a luminosity. Thus, the Firmani relation is less fundamental as it uses more parameters for the fit. Of the two relations that remain, the one that has the most `utility' will be the one with the least amount of scatter in its calibration curve. We were able to reproduce Firmani's results over his small sample of 19 bursts. However, when we substitute independent values for $L_{iso}$, $E_{peak}$ and $T_{0.45}$, we see a substantial broadening around the model. That is, the Firmani relation is not robust on the use of alternative input data. In addition, when we extend the test to a larger sample of 60 bursts, the scatter becomes substantially larger again. Indeed, this scatter is comparable to the scatter in the original $E_{peak} - L_{iso}$ relation. That is, the Firmani relation is not robust for the use of additional bursts. These failures of the Firmani relation have dashed our hopes raised by the tight calibration curves displayed in Firmani et al. (2006). It also suggests that the addition of a duration does not significantly improve the $L_{iso} - E_{peak}$ relation. The larger point of interest is that no duration shows a significant advantage over the $E_{peak} - L_{iso}$. While it might be possible that that a relation involving $\\mathcal{T}_{30}/N_{peak}$ might really have a smaller scatter than the Firmani relation, the improvements are small and not significant. This leads us to conclude that the addition of a duration is not doing enough to improve the $L_{iso} - E_{peak}$ relation to be considered to be a separate luminosity relation. We conclude that the Firmani relation is not useful for several reasons: First, the Firmani relation is simply derived by putting together two well-known, simpler, and independent luminosity relations, and thus it has no separate existence. Second, it is not robust for the inclusion of independent input data or for the extension to many more GRBs. Third, the real scatter for the Firmani relation does not live up to the hope generated by the original report, with the scatter being comparable to those of the luminosity relations from which it is derived. In all, we can see no utility or advantage to using the Firmani relation. We would like to thank the Louisiana State University Board of Regents and the Louisiana Space Consortium for their support." }, "0808/0808.2899_arXiv.txt": { "abstract": "We use $16\\;\\rm \\mu m$, Spitzer-IRS, blue peakup photometry of 50 early-type galaxies in the Coma cluster to define the mid-infrared colour-magnitude relation. We compare with recent simple stellar population models that include the mid-infrared emission from the extended, dusty envelopes of evolved stars. The K$_{\\rm s}$-[16] colour in these models is very sensitive to the relative population of dusty Asymptotic Giant Branch (AGB) stars. We find that the \\emph{passively evolving} early-type galaxies define a sequence of approximately constant age ($\\sim 10$~Gyr) with varying metallicity. Several galaxies that lie on the optical/near-infrared colour-magnitude relation do not lie on the mid-infrared relation. This illustrates the sensitivity of the K${\\rm s}$-[16] colour to age. The fact that a colour-magnitude relation is seen in the mid-infrared underlines the extremely passive nature of the majority (68 \\%) of early-type galaxies in the Coma cluster. The corollary of this is that 32\\% of the early-type galaxies in our sample are \\emph{not} `passive', insofar as they are either significantly younger than 10 Gyr or they have had some rejuvenation episode within the last few Gyr. ", "introduction": "Early-type galaxies (ellipticals and S0s, ETGs hereafter) are ancient stellar populations whose epoch of formation is related to the large-scale structure formation in the Universe. This picture is supported by various studies that have used optical line-strength indices to determine evolutionary parameters of ETGs in the cluster and field environments (see Renzini 2006 for a review). Among these studies, some authors (Bernardi et al. 2005, Clemens et al. 2006, Thomas et al. 2005, S{\\'a}nchez-Bl{\\'a}zquez et al. 2006a, Annibali et al. 2007) suggest that cluster ETGs have a luminosity weighted, mean stellar age 1-2 Gyr older than those in the field. Recently, Trager, Faber \\& Dressler (2008) challenged this view. Analyzing the Coma cluster, they found that the 12 ETGs in their sample have mean single stellar population equivalent ages of 5-8 Gyr, with the oldest systems being $\\leq$ 10 Gyr old. This average age is remarkably similar to the mean age of ETGs in low density environments (see e.g. Annibali et al. 2007 and references therein). \\begin{figure*} \\centerline{ \\includegraphics[scale=0.45]{31.ps} \\hskip-3mm \\includegraphics[scale=0.45]{49.ps} \\hskip-3mm \\includegraphics[scale=0.45]{78.ps} \\hskip-3mm \\includegraphics[scale=0.45]{95.ps} } \\vskip-5mm \\centerline{ \\includegraphics[scale=0.45]{129.ps} \\hskip-3mm \\includegraphics[scale=0.45]{143.ps} \\hskip-3mm \\includegraphics[scale=0.45]{144.ps} \\hskip-3mm \\includegraphics[scale=0.45]{148.ps} } \\vskip-5mm \\centerline{ \\includegraphics[scale=0.45]{167.ps} \\hskip-3mm \\includegraphics[scale=0.45]{172.ps} \\hskip-3mm \\includegraphics[scale=0.45]{240.ps} \\hskip-3mm \\includegraphics[scale=0.45]{psf.ps} } \\caption{Example background subtracted, PBCD, IRS peakup images of the Coma ETGs. Contour levels are $2^{n/2}/100$ where n=(-1,0,1,2\\dots) $\\rm mJy/pixel$ with a dashed contour for negative values at the lowest level. The grey-scale in each plot is scaled to the peak source flux. The last panel is an image of the peak-up point spread function. The cross in each panel gives the optical source position. The grey-scale has a square-root stretch and contour levels are $2^{n/2}/100$ where n=(1,3,5\\dots) $\\rm mJy/pixel$. Pixels are {1}\\farcs{8} on a side.} \\label{fig:maps1} \\end{figure*} The emission from elliptical galaxies longward of $\\sim 10\\;\\rm \\mu m$ has a large contribution from the circumstellar dust around evolved stars, such as those on the AGB. This dust emission has been detected by Spitzer in early-type galaxies where it is seen as a wide emission feature near $10\\;\\rm \\mu m$ with another broad feature near $18\\;\\rm \\mu m$ (Bressan et al., 2006). The spatial profile in the mid-infrared is similar to the optical profile (Temi et al., 2008, Athey et al. 2002), re-enforcing the notion that the origin is stellar rather than ISM. This cicumstellar dust has also been detected directly as an extended envelope around nearby AGB stars (Gledhill \\& Yates, 2003). Longer wavelength infrared emission from elliptical galaxies is sometimes detected (Leeuw et al., 2004; Marleau et al., 2006; Temi, Brighenti \\& Mathews, 2007.), but this cooler component is probably associated with a diffuse interstellar medium, rather than with evolved stars directly. For objects older than $0.1\\;\\rm Gyr$ the mid-infrared emission traces stars during their mass-losing AGB phase, and this evolutionary phase is still detected in objects with ages typical of globular clusters such as 47 Tucanae (Lebzelter et al., 2006). Because the luminosity of an AGB star depends on its mass and the main sequence turn-off mass decreases with time, the mid-infrared emission depends on the age of the stellar population. If all the stars in an early-type galaxy were created in an instantaneous burst $10\\;\\rm Gyr$ ago, the mid-infrared emission would be dominated by stars on the AGB with an initial mass of slightly less than $1\\;\\rm M_{\\odot}$ (for solar metallicity). The strength of the silicate emission from the dusty envelopes of evolved stars is a function of both age and metallicity. As pointed out by Bressan, Granato \\& Silva (1998) however, the dependence of this emission on age and metallicity is different to that in the optical. A colour-magnitude relation which includes a band containing the silicate emission from AGB stars therefore traces the AGB population as a function of the total luminosity, and can, in principle, be used with the optical relation to disentangle the effects of age and metallicity. Here, we use $16\\;\\rm \\mu m$ images of 50 ETGs in the Coma cluster, made with the blue peakup detector of the IRS on Spitzer. We combine these with archival Spitzer IRAC images at $4.5\\;\\rm\\mu m$ and K-band images to construct the mid-infrared colour-magnitude relation. As objects at the distance of the Coma cluster were too faint for IRS spectroscopy we include 4 galaxies in the Virgo cluster for which we have IRS spectra. These 4 objects were selected to have spectra that show no evidence of activity, such as recent star formation or an AGN, and are intended as a `template' for passive objects against which to compare the more distant objects of Coma. The Coma cluster is both very rich and dynamically relaxed, and contains some of the the most massive galaxies in the local Universe. As such, it is an ideal place to study the star formation history of ETGs. \\begin{figure} \\centerline{ \\includegraphics[scale=0.4]{coma_Re_log.ps} } \\caption{Relation between the effective H-band radii ($1.65\\;\\rm \\mu m$, Gavazzi et al. 2000) and the $16\\;\\rm \\mu m$ effective radii. The latter have been estimated by convolving a model $r^{1/4}$ de Vaucouleur's law with the IRS blue peakup PSF (see text). Crosses are ellipticals and diamonds are lenticulars (classified either as S0/E, S0, SB0, or S0/a). The solid line is a least squares fit to the data whereas the dotted line corresponds to equal Re at both wavelengths. The horizontal dot-dashed line corresponds to the radius containing half of the encircled energy for the IRS blue peakup PSF.} \\label{fig:radii} \\end{figure} ", "conclusions": "We present $16\\;\\rm \\mu m$, Spitzer-IRS, blue peakup images of a sample of 50 ETGs in the Coma cluster. We compare these with archival IRAC images at $4.5\\;\\rm \\mu m$ and 2MASS, $K_s$ band images at $2.2\\;\\rm \\mu m$. We make the following conclusions. \\begin{itemize} \\item{Our IRS blue peakup images show no evidence that the $16\\;\\rm \\mu m$ emission is anything other than stellar in origin} \\item{The region within a dusty AGB star envelope where dust is hot enough to emit strongly at $16\\;\\rm \\mu m$ is not vulnerable to environmental effects such as the ram-pressure of an ISM wind or dust sputtering by hot gas. Dust grains only spend $\\sim 100\\;\\rm yr$ near $\\sim 300\\;\\rm K$ and this region is sufficiently dense to survive against ram-pressure disruption.} \\item{We identify the mid-infrared colour-magnitude relation of passively evolving ETGs as the lower envelope of the galaxy distribution in the K$_s$ - [16] vs K$_s$ plane; that is, the minimum value of K$_s$ - [16] at a given K$_s$ magnitude. $\\sim$68\\% of the galaxies in our sample lie on this colour-magnitude relation. These galaxies cannot have had any episode of star formation accounting for more than $\\sim 1\\%$ of the total stellar mass within the last few Gyr. These are genuinely `passively evolving' objects. The remaining objects are either significantly younger than 10 Gyr or have undergone a rejuvenation event in the recent past. This result is at odds with the most recent estimate of the fraction of old objects based on optical spectroscopy (Trager et al. 2008).} \\item{We construct the mixed optical-NIR-MIR two colour diagram and, by means of updated simple stellar population models, we show that the addition of mid-infrared data allows a much better separation of the effects of age and metallicity, which are rather degenerate in either the optical or mid-infrared when taken in isolation. In this plane galaxies populating the colour-magnitude relation trace a sequence of \\emph{varying metallicity at approximately constant age}. Although this conclusion was already consistent with the optical-NIR colour-magnitude relation, a correlation between age and metallicity could equally well explain the relation. Indeed, comparison with the optical-NIR colours shows that \\emph{a number of galaxies that lie on the optical-NIR relation are significantly displaced from the mid-infrared relation}, with redder K$_s$ - [16] colours. The mid-infrared colour-magnitude diagram therefore shows a sequence of metallicity for old, passive galaxies, with younger objects displaced towards redder K$_s$ - [16] colours.} \\item{The oldest elliptical galaxies in our sample have luminosity weighted, mean stellar ages of 10.5 Gyr and metallicities within a factor of two of the solar value. No galaxy in our sample is as old or metal poor as the globular cluster 47 Tuc.} \\item{Although the addition of the K$_s$ - [16] colour allows us to identify objects with significantly younger, luminosity weighted, mean stellar ages, we cannot distinguish between genuinely `young' objects and those that have undergone a minor rejuvenation event. However, given that even a period of recent (last few Gyr) star formation that accounts for less than 1\\% of the total stellar mass will shift a galaxy off the mid-infrared colour-magnitude relation, the latter option seems far more likely. Unambiguous resolution of this issue will require the infrared spectroscopic capabilities of future space observatories.} \\item{There is evidence that those galaxies that have an excess in the K$_s$ - [16] colour are found preferentially at smaller cluster-centric radii. As the interaction between the dusty AGB star envelopes and the ISM does not directly effect the $16\\;\\rm \\mu m$ emission, the excess may be caused by ``rejuvenation'' episodes induced by the cluster environment.} \\end{itemize}" }, "0808/0808.3584_arXiv.txt": { "abstract": "We estimate cluster ages from lithium depletion in five pre-main-sequence groups found within 100\\,pc of the Sun: TW~Hydrae Association, $\\eta$~Chamaeleontis Cluster, $\\beta$~Pictoris Moving Group, Tucanae-Horologium Association and AB~Doradus Moving Group. We determine surface gravities, effective temperatures and lithium abundances for over 900 spectra through least squares fitting to model-atmosphere spectra. For each group, we compare the dependence of lithium abundance on temperature with isochrones from pre-main-sequence evolutionary tracks to obtain model dependent ages. We find that the $\\eta$\\,Chamaelontis Cluster and the TW~Hydrae Association are the youngest, with ages of $12{\\pm6}$\\,Myr and $12{\\pm8}$\\,Myr, respectively, followed by the $\\beta$~Pictoris Moving Group at $21{\\pm9}$\\,Myr, the Tucanae-Horologium Association at $27{\\pm11}$\\,Myr, and the AB Doradus Moving Group at an age of at least $45$\\,Myr (where we can only set a lower limit since the models -- unlike real stars -- do not show much lithium depletion beyond this age). Here, the ordering is robust, but the precise ages depend on our choice of both atmospheric and evolutionary models. As a result, while our ages are consistent with estimates based on Hertzsprung-Russell isochrone fitting and dynamical expansion, they are not yet more precise. Our observations do show that with improved models, much stronger constraints should be feasible: the intrinsic uncertainties, as measured from the scatter between measurements from different spectra of the same star, are very low: around 10\\,K in effective temperature, 0.05\\,dex in surface gravity, and 0.03\\,dex in lithium abundance. ", "introduction": "\\label{s:intro} Triggered largely by the discovery of young stars in the {\\em ROSAT} X-ray Satellite All-Sky Survey, over the last decade several nearby pre-main-sequence (PMS) star groups have been identified (for a review, see \\citealt{zuc04}). Ranging in age from roughly 6\\,Myr to $\\sim$100\\,Myr, five separate groups can be distinguished: TW Hydrae Association (TWA), $\\eta$~Chamaeleontis cluster ($\\eta$\\,Cha), $\\beta$~Pictoris Moving Group (BPMG), Tucanae-Horologium association (TUCHOR) and AB~Doradus Moving Group (ABD). The common space motions and localized sky positions suggest that these groups are likely connected to the Sco-Cen star forming region located $\\sim$100\\,pc away in the southern hemisphere. \\citet{mam99} and \\citet{son03} have traced back the space motion of members in BPMG, TWA and $\\eta$\\,Cha and argue that the groups are related to a star formation burst in the Sco-Cen region as a result of the passing of the Carina arm $\\sim$60\\,Myr ago. These groups, because of their close vicinity, are excellent laboratories for studying star and planet formation. Well constrained ages are necessary to make conclusions about timescales of, e.g., disk dissipation and planet formation. Already, observations from the same sample presented in this paper have revealed that accretion disks can last up to $\\sim$10\\,Myr, but beyond this is rare \\citep{jay06}. \\defcitealias{bar98}{BCAH98} Previously, ages have been derived from Hertzsprung-Russell (HR) diagram fitting, group dynamics and lithium abundance measurements. \\citet{luh04} provide an HR diagram isochrone age for $\\eta$\\,Cha of $6^{+2}_{-1}$\\,Myr derived from the evolutionary models of \\citeauthor{bar98} (\\citeyear{bar98}; hereafter \\citetalias{bar98}) and \\citet{pal99}, which agrees well with the dynamical expansion age of 6.7\\,Myr determined by \\citet{jil05}. The dynamical age of TWA has been harder to determine because of inconsistent space motions among its more than 30 members. An inferred age of $8.3\\pm0.8$\\,Myr is given to TWA based on the dynamical motion of four members \\citep{del06}. However, the likely complex dynamical evolution of TWA has led to several plausible evolutionary scenarios. \\citet{mak05} attribute this complex evolution to a chance encounter with Vega, while \\citet{law05} suggest TWA is composed of two separate groups, based on bimodal rotation period distributions with distinctly separate ages of $\\sim$\\,10\\,Myr and $\\sim$\\,17\\,Myr. More recently, \\citet{bar06} finds a conservative age of $10^{+10}_{-7}$\\,Myr by comparing ages from HR diagram isochrone comparisons (from BCAH98) and lithium abundances. The slightly older group BPMG has an estimated age of $12^{+8}_{-4}$\\,Myr based on HR diagram isochrone comparisons (from \\citetalias{bar98}) and lithium abundances \\citep{zuc01}, with three dimensional motions that are consistent with a dynamical expansion age of 11.5\\,Myr \\citep{ort02}. \\citet{fei06} independently derived an age of $13^{+4}_{-3}$\\,Myr for the recently confirmed wide binary system of 51~Eri and GJ\\,3305, part of BPMG. Known to be older than BPMG, but younger than the Pleiades, TUCHOR has an age of 20--40\\,Myr based on H$\\alpha$ measurements, X-ray luminosity, rotation and lithium abundances in comparison to other young clusters like TWA and the Pleiades \\citep{zuc00,ste00}. Perhaps, the most debated age is that of the ABD group. \\citet{zuc04b} derived an age of 50\\,Myr by comparing the H$\\alpha$ emission strength of ABD members to members of the younger TUCHOR association, in addition to fitting its three M-type members to HR diagram isochrones. In contrast, \\citet{luh05} compared HR isochrones of ABD members to those of two well-observed clusters with ages of 50 and 125\\,Myr and suggested that the ABD group is coeval with the Pleiades at an age of 100--125\\,Myr. The latter age is strongly supported by \\citet{ort07}, who compute full 3D galactic orbits of ABD and the Pleiades cluster and show the dynamics of the two groups can be traced back to a common origin of $119\\pm20$\\,Myr ago. A relatively new approach to age estimates is to use the evolution of the lithium abundance for low-mass, partially and fully convective PMS stars \\citep{bil97,jef05}. The initiation and duration of lithium depletion in PMS stars is dependent on mass and is very sensitive to the central temperature. Lithium is converted into helium in $p,\\alpha$ reactions in cores of low-mass stars when the temperature reaches $2.5\\times10^{6}$\\,K. The lower the stellar mass, the longer the time it takes to reach this critical temperature. For example, a 0.6\\,\\Msun\\ star begins to burn lithium at an age of 3\\,Myr, while a lower mass star at 0.1\\,\\Msun\\ begins to burn lithium at an age of 40\\,Myr. Stars with $M<0.06$\\,\\Msun\\ never reach this typical temperature, while stars with 0.6\\,\\Msun\\ $2$. This may indicate that there is a time delay between SF activity and QSO activity, which is naturally explained in the merger picture (any galaxy with cold gas can form stars, but galaxies have to ``wait'' for a major merger before significant accretion onto the BH can occur). One way to test this picture, in which mergers are responsible for triggering both AGN activity and morphological and spectrophotometric transformation of galaxies, is to build cosmological models that treat the growth of galaxies, black holes, and AGN self-consistently. In addition, many astronomers now believe that the energy released by accreting black holes may play a crucial role in regulating galaxy formation. Here we describe one such semi-analytic model for the joint formation of galaxies, black holes, and AGN, and present some predictions from these models. ", "conclusions": "" }, "0808/0808.2055.txt": { "abstract": "In this paper, we present a systematical study of brane worlds of string theory on $S^{1}/Z_{2}$. In particular, starting with the toroidal compactification of the Neveu-Schwarz/Neveu-Schwarz sector in (D+d) dimensions, we first obtain an effective $D$-dimensional action, and then compactify one of the $(D-1)$ spatial dimensions by introducing two orbifold branes as its boundaries. We divide the whole set of the gravitational and matter field equations into two groups, one holds outside the two branes, and the other holds on them. By combining the Gauss-Codacci and Lanczos equations, we write down explicitly the general gravitational field equations on each of the two branes, while using distribution theory we express the matter field equations on the branes in terms of the discontinuities of the first derivatives of the matter fields. Afterwards, we address three important issues: (i) the hierarchy problem; (ii) the radion mass; and (iii) the localization of gravity, the 4-dimensional Newtonian effective potential and the Yukawa corrections due to the gravitational high-order Kaluza-Klein (KK) modes. The mechanism of solving the hierarchy problem is essentially the combination of the large extra dimension and warped factor mechanisms together with the tension coupling scenario. With very conservative arguments, we find that the radion mass is of the order of $10^{-2}\\; GeV$. The gravity is localized on the visible brane, and the spectrum of the gravitational KK modes is discrete and can be of the order of TeV. The corrections to the 4-dimensional Newtonian potential from the higher order of gravitational KK modes are exponentially suppressed and can be safely neglected in current experiments. In an appendix, we also present a systematical and pedagogical study of the Gauss-Codacci equations and Israel's junction conditions across a (D-1)-dimensional hypersurface, which can be either spacelike or timelike. ", "introduction": "\\renewcommand{\\theequation}{1.\\arabic{equation}} \\setcounter{equation}{0} Superstring and M-theory all suggest that we may live in a world that has more than three spatial dimensions.\u00a0 Because only three of these are presently observable, one has to explain why the others are hidden from detection.\u00a0 One such explanation is the so-called Kaluza-Klein (KK) compactification, according to which the size of the extra dimensions is very small (often taken to be on the order of the Planck length).\u00a0 As a consequence, modes that have momentum in the directions of the extra dimensions are excited at currently inaccessible energies. Recently, the braneworld scenarios \\cite{ADD98,RS1} has dramatically changed this point of view and, in the process, received a great deal of attention. At present, there are a number of proposed models (See, for example, \\cite{branes} and references therein.). In particular, Arkani-Hamed {\\em et al} (ADD) \\cite{ADD98} pointed out that the extra dimensions need not necessarily be small and may even be on the scale of millimeters.\u00a0 This model assumes that Standard Model fields are confined to a three (spatial) dimensional hypersurface (a 3-brane) living in a larger dimensional bulk while the gravitational field propagates in the bulk.\u00a0 Additional fields may live only on the brane or in the bulk, provided that their current undetectability is consistent with experimental bounds. One of the most attractive features of this model is that it may potentially resolve the long standing hierarchy problem, namely the large difference in magnitudes between the Planck and electroweak scales, %\\bq %\\lb{1.0} %\\frac ${M_{pl}}/{M_{EW}} \\simeq 10^{16}$, %\\eq where $M_{pl}$ denotes the four-dimensional Planck mass with $M_{pl} \\sim 10^{16} \\; TeV$, and $M_{EW}$ the electroweak scale with $M_{EW} \\sim TeV$. Considering a N-dimensional spacetime and assuming that the extra dimensions are homogeneous and finite, we find \\bqn \\label{1.1} S^{(N)}_{g} &\\sim& -{M^{N-2}} \\int{dx^{4} dz^{n}\\sqrt{- g^{(N)}} R^{(N)}} \\nb\\\\ &=& -{M^{N-2}V_{n}}\\int{d^{4}x \\sqrt{- g^{(4)}} R^{(4)}}\\nb\\\\ &\\simeq& -{M^{2}_{pl} }\\int{d^{4}x \\sqrt{- g^{(4)}} R^{(4)}}, \\eqn where $V_{n}$ denotes the volume of extra dimensions, $n \\equiv N-4$, and $M$ the N-dimensional fundamental Planck mass, which is related to $M_{pl}$ by \\bq \\lb{1.2} M = \\left({M^{2}_{pl}}/{V_{n}}\\right)^{1/(2+n)}. \\eq Clearly, for any given extra dimensions, if $V_{n}$ is large enough, $M$ can be as low as the electroweak scale, $M \\simeq M_{EW} \\simeq TeV$. Therefore, if we consider $M$ as the fundamental scale and $M_{pl}$ the deduced one, we can see that the hierarchy between the two scales is exactly due to the dilution of the spacetime in high dimensions, whereby the hierarchy problem is resolved. Table top experiments show that Newtonian gravity is valid at least down to the size $R \\sim 44 \\; \\mu{m}$ \\cite{Hoyle04}. From the above we can see that for $n \\ge 2$ the N-dimensional Planck mass $M$ can be lowered down to the electroweak scale from the four-dimensional Planck scale. In a different model, Randall and Sundrum (RS1) \\cite{RS1} showed that if the self-gravity of the brane is included, gravitational effects can be localized near the Planck brane at low energy and the 4D Newtonian gravity is reproduced on this brane. In this model, the extra dimensions are not homogeneous, but warped. One of the most attractive features of the model is that it will soon be fully explored by LHC \\cite{DHR00}. In the RS1 scenario, the mechanism to solve the hierarchy problem is completely different \\cite{RS1}. Instead of using large dimensions, RS used the warped factor, for which the mass $m_{0}$ measured on the invisible (Planck) brane is related to the mass $m$ measured on the visible (TeV) brane by $m = e^{-ky_{c}}m_{0}$, where $e^{-ky_{c}}$ is the warped factor. Clearly, by properly choosing the distance $y_{c}$ between the two branes, one can lower $m$ to the order of $TeV$, even $m_{0}$ is of the order of $M_{pl}$. It should be noted that the five-dimensional Planck mass $M_{5}$ in the RS1 scenario is still in the same order of $M_{pl}$. In fact, the 5-dimensional action $S^{5)}_{g}$ can be written as \\bqn \\lb{1.3} S^{(5)}_{g} &\\sim& -{M^{3}}\\int{dx^{4} d\\phi\\sqrt{- g^{(5)}} R^{(5)}}\\nb\\\\ &=& -{M^{3}}\\int^{\\pi}_{-\\pi}{r_{c}e^{-2kr_{c}|\\phi|}d\\phi} \\int{d^{4}x \\sqrt{- g^{(4)}} R^{(4)}} \\nb\\\\ &\\simeq& -{M^{2}_{pl} }\\int{d^{4}x \\sqrt{- g^{(4)}} R^{(4)}}, \\eqn where now we have \\bq \\lb{1.4} M^{2}_{pl} = {M^{3}}{k^{-1}}\\left(1 - e^{-2ky_{c}}\\right) \\simeq M^{2}_{5}, \\eq for $k \\simeq M_{5}$ and $ky_{c} \\simeq 35$. Another long-standing problem is the cosmological constant problem: Its theoretical expectation values from quantum field theory exceed its observational limits by $120$ orders of magnitude \\cite{wen}. Even if such high energies are suppressed by supersymmetry, the electroweak corrections are still $56$ orders higher. This problem was further sharpened by recent observations of supernova (SN) Ia, which reveal the revolutionary discovery that our universe has lately been in its accelerated expansion phase \\cite{agr98}. Cross checks from the cosmic microwave background radiation and large scale structure all confirm it \\cite{Obs}. In Einstein's theory of gravity, such an expansion can be achieved by a tiny positive cosmological constant, which is well consistent with all observations carried out so far \\cite{SCH07}. Because of this remarkable result, a large number of ambitious projects have been aimed to distinguish the cosmological constant from dynamical dark energy models \\cite{DETF}. Since the problem is intimately related to quantum gravity, its solution is expected to come from quantum gravity, too. At the present, string/M-Theory is our best bet for a consistent quantum theory of gravity, so it is reasonable to ask what string/M-Theory has to say about the cosmological constant. In the string landscape \\cite{Susk}, it is expected there are many different vacua with different local cosmological constants \\cite{BP00}. Using the anthropic principle, one may select the low energy vacuum in which we can exist. However, many theorists still hope to explain the problem without invoking the existence of ourselves. In addition, to have a late time accelerating universe from string/M-Theory, Townsend and Wohlfarth \\cite{townsend} invoked a time-dependent compactification of pure gravity in higher dimensions with hyperbolic internal space to circumvent Gibbons' non-go theorem \\cite{gibbons}. Their exact solution exhibits a short period of acceleration. The solution is the zero-flux limit of spacelike branes \\cite{ohta}. If non-zero flux or forms are turned on, a transient acceleration exists for both compact internal hyperbolic and flat spaces \\cite{wohlfarth}. Other accelerating solutions by compactifying more complicated time-dependent internal spaces can be found in \\cite{string}. Recently, we studied brane cosmology in the framework of both string theory \\cite{WS07,WSVW08} and the Horava-Witten (HW) heterotic M Theory \\cite{GWW07,WGW08} on $S^{1}/Z_{2}$. From a pure numerology point of view, we found that the 4D effective cosmological constant can be cast in the form, \\bq \\lb{1.5} \\rho_{\\Lambda} = \\frac{\\Lambda_{4}}{8\\pi G_{4}} = 3\\left(\\frac{R}{l_{pl}}\\right)^{\\alpha_{R}}\\left(\\frac{M}{M_{pl}}\\right)^{\\alpha_{M}} M_{pl}^{4}, \\eq where $R$ denotes the typical size of the extra dimensions, $M$ is the energy scale of string or M theory, and $(\\alpha_{R}, \\alpha_{M}) = (10, 16)$ for string theory and $(\\alpha_{R}, \\alpha_{M}) = (12, 18)$ for the HW heterotic M Theory. In both cases, it can be shown that for $R \\simeq 10^{-22} \\; m$ and $M_{10} \\simeq 1\\; TeV$, we obtain $\\rho_{\\Lambda} \\sim \\rho_{\\Lambda, ob} \\simeq 10^{-47}\\; GeV^{4}$. When orbifold branes are concerned, a critical ingredient is the radion stability. Using the mechanism of Goldberger and Wise \\cite{GW99}, we showed that the radion is stable. Such studies were also generalized to the HW heterotic M Theory \\cite{HW96}, and found that, among other things, the radion is stable and has a mass of order of $10^{-2} \\; GeV$ \\cite{WGW08}. In this paper, we shall give a systematical study of brane worlds of string theory on $S^{1}/Z_{2}$. Similar studies in 5-dimensional spacetimes have been carried out in the framework of both string theory \\cite{WSVW08} and M Theory \\cite{WGW08}. However, to have this paper as much independent as possible, it is difficult to avoid repeating some of our previous materials, although we would try our best to keep it to its minimum. The rest of the paper is organized as follows: In Sec. II, we consider the toroidal compactification of the Neveu-Schwarz/Neveu-Schwarz (NS-NS) sector in (D+d) dimensions, and obtain an effective $D$-dimensional action. Then, we compactify one of the $(D-1)$ spatial dimensions by introducing two orbifold branes as the boundaries along this compactified dimension. In Sec. III, we divide the whole set of the gravitational and matter field equations into two groups, one holds outside the two branes, and the other holds on each of them. Combining the Gauss-Codacci and Lanczos equations, we write down explicitly the general gravitational field equations on the branes, while using distribution theory we are able to express the matter field equations on the branes in terms of the discontinuities of the first derivatives of the matter fields. In Sec. IV, we study the hierarchy problem, while in Sec. V, we consider the radion mass by using the Goldberger-Wise mechanism \\cite{GW99}. In Sec. VI we study the localization of gravity, the 4-dimensional effective potential and high order Yukawa corrections. In Sec. VII, we present our main conclusions with some discussing remarks. We also include an appendix, in which we present a systematical and pedagogical study of the Gauss-Codacci equations and Israel's junction conditions across a surface, where the metric coefficients are only continuous, i.e., $C^{0}$, in higher dimensional spacetimes. To keep such a treatment as general as possible, the surface can be either spacelike or timelike. Before turning to the next section, we would like to note that in 4-dimensional spacetimes there exists Weinberg's no-go theorem for the adjustment of the cosmological constant \\cite{wen}. However, in higher dimensional spacetimes, the 4-dimensional vacuum energy on the brane does not necessarily give rise to an effective 4-dimensional cosmological constant. Instead, it may only curve the bulk, while leaving the brane still flat \\cite{CEG01}, whereby Weinberg's no-go theorem is evaded. It was exactly in this vein, the cosmological constant problem was studied in the framework of brane worlds in 5-dimensional spacetimes \\cite{5CC} and 6-dimensional supergravity \\cite{6CC}. However, it was soon found that in the 5-dimensional case hidden fine-tunings are required \\cite{For00}. In the 6-dimensional case such fine-tunings may not be needed, but it is still not clear whether loop corrections can be as small as expected \\cite{Burg07}. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "\\renewcommand{\\theequation}{7.\\arabic{equation}} \\setcounter{equation}{0} In this paper, we have systematically studied the brane worlds of string theory on $S^{1}/Z_{2}$. Starting with the toroidal compactification of the Neveu-Schwarz/Neveu-Schwarz sector in (D+d) dimensions, in Sec. II.A we have first obtained an effective $D$-dimensional action given by Eq.(\\ref{2.16}) for non-vanishing dilaton field and flux with an effective potential given by Eq.(\\ref{2.17}). Then, in Sec. II.B we have compactified one of the $(D-1)$ spatial dimensions by adding two orbifold branes as the boundaries of the spacetime along the compactified dimension. Variations of the total action with the metric and matter fields yield, respectively, the gravitational and matter field equations. This has been done in Sec. III and given by Eqs.(\\ref{3.3})-(\\ref{3.ee}). Dividing the whole set of the field equations into two groups, one holds outside the two branes, and the other holds on them, in Sec. III.A we have first written down the field equations outside the two branes, Eqs.(\\ref{3.ef})-(\\ref{3.ej}), while in Sec. III.B, we have written down explicitly the general gravitational field equations on each of the two branes, Eqs. (\\ref{3.15})-(\\ref{3.17}), by combining the Gauss-Codacci and Lanczos equations. On the other hand, by using the distribution theory, we have also been able to write down the matter field equations on the branes in terms of the discontinuities of the first derivatives of the matter fields, Eqs. (\\ref{3.27a})-(\\ref{3.28}). In the study of orbifold branes, one of the most attractive features is that it may resolve the long standing hierarchy problem. In Sec. IV, we have shown explicitly how it can be solved in the current setup. The mechanism is essentially the combination of the ADD large extra dimension \\cite{ADD98} and RS warped factor \\cite{RS1} mechanisms together with the tension coupling scenario \\cite{Cline99}. In order to solve the hierarchy problem in the current setup, the tensions of the branes are required to be in the order of the cosmological constant. %, which might represent fine-tuning. %It should be noted that this result %is quite general, and applicable to a large class of brane-world scenarios %\\cite{branes}. Another important issue in brane worlds is the radion stability and radion mass \\cite{branes}. Previously, we showed that the radion is stable \\cite{WS07}. In this paper, we have devoted Sec. V to study the radion mass. With some very conservative arguments, we have found that the radion mass is of the order of $10^{-2}\\; GeV$, which is by far beyond its current observational constraint, $m_{\\varphi} > 10^{-3}\\; eV$. In Sec. VI we have also shown that the gravity is localized on the visible (TeV) brane, in contrast to the RS1 model in which the gravity is localized on the Planck (hidden) brane \\cite{RS1}. In addition, the spectrum of the gravitational KK modes is discrete, and given explicitly by Eq.(\\ref{7.23}), which can be of the order of TeV. The corrections to the 4D Newtonian potential from the higher order gravitational KK modes are exponentially suppressed and can be safely neglected [cf. Eq.(\\ref{7.27})]. In Appendix, we have also presented a systematical and pedagogical study of the Gauss-Codacci equations and Israel's junction conditions across a surface, which can be either spacelike or timelike, in higher dimensional spacetimes. It should be noted that, when studied the radion stability, we have ignored the backreaction of the perturbations. Although it is expected that the main results obtained here will be continuously valid even after taking such backreaction into acocunt, as what exactly happened in the Randall-Sundrum model \\cite{radion}, it would be very interesting to show explicitly that this is indeed the case. Other important issues that have not been addressed in this paper include the constraints from the solar system tests \\cite{solar}, and linear perturbations in the current setup." }, "0808/0808.0702_arXiv.txt": { "abstract": "\\noindent A number of positive and null results on the time variation of fundamental constants have been reported. It is difficult to judge whether or not these claims are mutually consistent, since the observable quantities depend on several parameters, namely the coupling strengths and masses of particles. The evolution of these coupling-parameters over cosmological history is also a priori unknown. A direct comparison requires a relation between the couplings. We explore several distinct scenarios based on unification of gauge couplings, providing a representative (though not exhaustive) sample of such relations. For each scenario we obtain a characteristic time dependence and discuss whether a monotonic time evolution is allowed. For all scenarios, some contradictions between different observations appear. We show how a clear observational determination of non-zero variations would test the dominant mechanism of varying couplings within unified theories. ", "introduction": "Any observation of a time variation of ``fundamental constants'' would be a far-reaching discovery. There are various claims for a detection, and many more observations indicating a null variation: criteria for judging their mutual consistency would be useful. We investigate whether a simple scenario exists which can account for several observational claims simultaneously and is consistent with unification of Standard Model (SM) gauge couplings. % Several observations motivate such a study. First, the claimed deviations from the present value of the fine structure constant $\\alpha$, or the proton-to-electron mass ratio $\\mu$, observed in quasar absorption systems. Second, the discrepancy between the primordial \\lise\\ abundance expected from standard nucleosynthesis (BBN) and seen in old halo stars, which may be explained by a variation of ``constants'' within a unified framework \\cite{DSW07,CocNunes06}. Third, the theoretical insight that scalar fields cannot be exactly constant over the entire cosmological evolution, and that a possible ``late'' time evolution can play an important role in the dynamics of the expanding Universe. A time variation of couplings arising from the evolution of a ``cosmon'' field in so-called ``quintessence scenarios'' \\cite{Wetterich:1987fk,DvaliZ} would link these variations to observables in cosmology \\cite{Wetterich:2002ic,Nunes:2003ff}. Due to the many unknowns of the underlying particle physics models and the partly contradictory present observational situation, a systematic treatment is not easy, and may even seem premature. Nevertheless we consider a first attempt to be useful, in order to discuss strategies that can be used to compare variations of different observables. The power of the proposed method will only become clear if and when future observations present a less ambiguous picture. The basic approach in the present paper relates the fractional variations of different fundamental couplings $G_k$, such as the fine structure constant $\\alpha$, the proton-electron mass ratio $\\mu$ or the ratio of the nucleon mass to the Planck mass, by an assumption of {\\it proportionality}, with fixed ``unification coefficients'' $d_k$. The choice of the values of $d_k$ is in turn determined within different scenarios of varying parameters in unified theories (GUTs) where the gauge couplings of the Standard Model converge at a unification scale $M_X$. The assumption of time-independent coefficients $d_k$ covers a large class of possible models for varying couplings. This assumption is, however, not a necessity, and we will describe specific quintessence models where it may not be realized in a forthcoming paper \\cite{Part2}. In Section~\\ref{sec:Data} of this paper we review observational determinations of the variation or constancy of couplings, considering five types of methods: early-universe cosmology, astrophysical spectroscopy, nuclear physics in the Earth and the Solar System, gravitational physics, and atomic clock comparisons. In Section~\\ref{sec:Unification} we introduce the unification of couplings (Grand Unified Theory, GUT) and determine the implications of unification and supersymmetry (SUSY) for the Standard Model parameters and for the observables we consider. We further define six unified scenarios by considering different possibilities for the variation of the Fermi scale and the superpartner masses. Within each scenario we reduce the various observational results to constraints on the time evolution of a single fundamental coupling, and discuss the mutual consistency of observations. In Section~\\ref{sec:Epochs_and_evolutionFactors} six cosmological epochs are introduced, and the constraints on variation deduced in Section~\\ref{sec:Unification} are collected into a set of ``evolution factors'' for each unified scenario. These evolution factors are a measure of the overall size of coupling evolution between a given epoch and the present. We then determine for each scenario to what extent a monotonic variation over time can be consistent with the data. Section~\\ref{sec:Conclude} draws some general conclusions. In a subsequent paper \\cite{Part2} we will investigate the scalar field dynamics that could give rise to a small but nonzero variation of couplings. The presence of a cosmologically varying degree of freedom gives rise to important additional effects. It affects gravity on large scales, altering the expansion of the Universe and potentially giving rise to the observed late-time acceleration. Also on local scales, a light field weakly coupled to matter produces long-range forces which are tightly constrained \\cite{Will} by Solar System precision tests of gravity and the null results of experiments testing the Weak Equivalence Principle. Combining all these considerations leads to tighter constraints on models but also offers more possibilities to test them. ", "conclusions": "\\label{sec:Conclude} Within Grand Unified Theories, different measurements of the variation of fundamental constants can be consistently reduced to a variation of a few ``unification parameters'', namely the unification scale $M_X/M_{\\rm P}$, gauge coupling $\\alpha_X$, the Fermi scale $\\vev{\\phi}/M_X$ and SUSY-breaking masses $\\tilde{m}/M_X$. We define various GUT-scenarios for varying couplings by the assumption of proportionality of fractional variations of the unification parameters. Assuming that couplings really vary, this is a way of excluding such GUT scenarios by demanding consistency of the implied variations. The assumption of proportionality permits us to project all observations into constraints on a common evolution factor $l(z)$ for each scenario. We show that different GUT scenarios yield different time evolutions of $l(z)$ assuming that certain claimed measurements of varying constants are correct. We confirm that ``simple'' models which have only one fundamental parameter varying ($\\alpha_X$ or $\\vev{\\phi}/M_X$) result in inconsistent variations. However, combined variations of these two parameters, as described in scenarios 5 and 6, lead to results more consistent with the possible quintessence-induced time variations of fundamental couplings which we investigate in \\cite{Part2}. % Specifically, one may ask whether the claimed observations of variations in $\\alpha$ \\cite{Murphy:2003mi} and $\\mu$ \\cite{Reinhold:2006zn} are mutually consistent, and whether they are consistent with an explanation of the apparent primordial \\lise-depletion by varying couplings. Within a hypothesis of constant Yukawa couplings, which results in identical fractional variations of all quark and lepton masses, we investigated arbitrary variations of $\\alpha_X$, $\\vev{\\phi}/M_X$ and $M_X/M_{\\rm P}$. For scenarios with supersymmetry we also assumed that the SUSY-breaking masses vary proportional to $\\vev{\\phi}$, but the effect of such a variation only appears at higher order and is probably not crucial. We have not found a scenario with a monotonic time evolution $l(z)$ that makes all three signals or hints of variation mutually consistent. A monotonic evolution requires either to discount one of the ``signals'' by substantially increasing its uncertainty, or to alter our assumptions by including additional time variation of some Yukawa couplings. Our investigation shows how the variations of different couplings in the Standard Model may be compared. If the observational situation becomes clearer and at least one nonzero time variation is established, such methods may be used for new tests of the idea of grand unification. \\subsection*{Note added} Shortly before the completion of this paper a new determination of the variation of $\\mu$ appeared \\cite{King:2008ud} reporting a reanalysis of spectra from the same two H$_2$ absorption systems as \\cite{Reinhold:2006zn}, and adding one additional system at $z\\simeq 2.8$. The results of the new analysis are not consistent with the previous claim indicating a nonzero variation, either considering all three systems or the two previously considered. The stringent null bound of the new analysis, $\\Delta \\mu/\\mu = (2.6 \\pm 3.0) \\times 10^{-6}$, would disfavour all scenarios except those where the fractional variation of $\\mu$ was of the same order as or smaller than that of $\\alpha$. This would require us to approach the ``special'', apparently fine-tuned values of $\\tilde{\\gamma}$ discussed in Section~\\ref{sec:Specials}, for which $\\mu$ variation (and any deviation from the standard \\lise\\ abundance at BBN) are suppressed. \\subsection*{Acknowledgements} We acknowledge useful discussions with M.~Pospelov, R.~Trotta, P.~Molaro, P.~Avelino, J.~Berengut and V.~Flambaum, and invaluable correspondence and discussions with M.~Murphy. T.\\,D. is supported by the {\\em Impuls- and Vernetzungsfond der Helmholtz-Gesellschaft}. \\appendix \\newcommand{\\appsection}[1]{\\let\\oldthesection\\thesection \\renewcommand{\\thesection}{Appendix \\oldthesection}" }, "0808/0808.2641_arXiv.txt": { "abstract": "We study the potential of GLAST to unveil particle dark matter properties with gamma-ray observations of the Galactic center region. We present full GLAST simulations including all gamma-ray sources known to date in a region of 4 degrees around the Galactic center, in addition to the diffuse gamma-ray background and to the dark matter signal. We introduce DMFIT, a tool that allows one to fit gamma-ray emission from pair-annihilation of generic particle dark matter models and to extract information on the mass, normalization and annihilation branching ratios into Standard Model final states. We assess the impact and systematic effects of background modeling and theoretical priors on the reconstruction of dark matter particle properties. Our detailed simulations demonstrate that for some well motivated supersymmetric dark matter setups with one year of GLAST data it will be possible not only to significantly detect a dark matter signal over background, but also to estimate the dark matter mass and its dominant pair-annihilation mode. ", "introduction": "The Gamma-ray Large Area Space Telescope (GLAST) \\cite{glastref} was successfully launched on June 11, 2008. The main instrument onboard GLAST, the Large Area Telescope (LAT) represents an improvement by more than one order of magnitude over the sensitivity of its predecessor EGRET \\cite{egretref} in the energy range from 20 MeV to 10 GeV, and extends the high energy coverage to about 300 GeV. These features make the LAT a tremendous tool for the indirect search for particle dark matter \\cite{Baltz:2008wd}, expected in the best motivated particle models to be a weakly interacting massive particle (WIMP) with a mass in the 10-1000 GeV range \\cite{dmreviews,kkdm}. WIMPs occasionally pair-annihilate in dark matter halos, producing, as a result, Standard Model particles. Prompt production as well as decays, hadronization and radiative processes associated to the annihilation products give rise to stable species including gamma rays, in an energy range extending up to the WIMP mass. Secondary radiation from the energetic electrons and positrons produced in annihilation events can also be used as an indirect detection diagnostic, e.g. in X-ray and radio wavelengths \\cite{multiw}. The signal from dark matter annihilation is proportional to the product of the particle pair-annihilation rate and to the particle number density squared along the line of sight. The latter quantity provides a guideline as to which regions of the gamma-ray sky are the most promising for the detection of a signal from dark matter annihilation. As discussed e.g. in \\cite{Baltz:2008wd}, targets include the center of our own Galaxy, satellite galaxies of the Milky Way, nearby massive galaxies and clusters, as well as diffuse emission from the entire Galactic halo and the summed signal from unresolved sources at all redshifts in the extra-galactic diffuse gamma-ray flux. Numerous studies have addressed in detail all of these possibilities (for reviews, see e.g. \\cite{dmreviews,kkdm}). With the dawn of the GLAST era, theoretical speculations on indirect WIMP detection with gamma rays will soon face high-quality data and the challenge of the corresponding analysis. The main purpose of the present study is to introduce DMFIT, a numerical package that, interfaced with any spectral fitting routine, allows one to fit gamma-ray data with the expected emission from the annihilation of generic WIMP models. For the purpose of illustration, we use here DMFIT in conjunction with XSPEC \\cite{xspec}. We show fits to gamma-ray spectra simulated using the experimental response function of the LAT and the software analysis tools that will be used for the actual GLAST data analysis. We apply DMFIT to the case of an ``isolated'' gamma-ray source, such as might be associated to dark matter annihilation in Galactic dark matter clumps or local satellite galaxies, as well as to the complex case of the Galactic center region, where background modeling and an accurate understanding of conventional gamma-ray sources play a critical role. The Galactic center region has long been considered a promising target to look for a signature from particle dark matter annihilation in gamma rays. First estimates of the detectability of a dark matter pair-annihilation signal from the center of the Galaxy date back 30 years \\cite{Gunn:1978gr,Stecker:1978du}. An incomplete list of recent references that discussed the detection of dark matter pair-annihilation in the Galactic center with gamma-ray observations includes \\cite{Cesarini:2003nr,Stoehr:2003hf,Aloisio:2004hy,Bloom:2004aq,Pieri:2005vp,Mambrini:2005vk,Profumo:2005xd,Jacholkowska:2005nz,Hall:2006na,Zaharijas:2006qb,Bergstrom:2006ny,Aharonian:2006wh,Ahn:2007ty,Morselli:2007ch,Hooper:2007gi,Dodelson:2007gd,Regis:2008ij,Serpico:2008ga,Baltz:2008wd}. In particular, ref.~\\cite{Cesarini:2003nr} discussed the dark matter interpretation of the gamma-ray excess reported by EGRET from the Galactic center \\cite{MayerHasselwander:1998hg}; ref.~\\cite{Profumo:2005xd} and \\cite{Aharonian:2006wh} discussed a similar possibility for the high energy gamma-ray flux detected by HESS \\cite{Aharonian:2004wa}; ref.~\\cite{Zaharijas:2006qb,Dodelson:2007gd} presented studies where both the EGRET and the HESS data were simultaneously taken into account as a background to dark matter searches. In addition, recently, the GLAST collaboration gave in ref.~\\cite{Baltz:2008wd} a comprehensive and updated overview of the GLAST sensitivity to dark matter annihilation signals for several possible sources inside and outside the Galaxy, using the Collaboration's current state of the art Monte Carlo and event reconstruction software. As far as the Galactic center region is concerned, in the present study we include all gamma-ray sources known to-date within an angle of 4 degrees from the Galactic center, in addition to the diffuse Galactic gamma-ray emission. Of special importance is the question of how to properly model the innermost sources, associated to the Sag A${^*}$ region: we model those with three different scenarios, described in detail in sec.~\\ref{sec:saga}. We introduce three reference particle dark matter setups, which are particularly illustrative for the scope of the present analysis, chosen to be theoretically well motivated, phenomenologically viable and producing the correct thermal relic neutralino abundance. We carry out complete one year GLAST simulations, making use of an up-to-date LAT instrumental response function and pointing mode setup, as well as of the software analysis tools that will be used to analyze the actual GLAST data. The main scopes of the present study are to: \\begin{enumerate} \\item present the DMFIT tool, and show examples of its application; \\item provide an updated template for the gamma-ray sources potentially relevant to dark matter searches in the Galactic center region; \\item assess the capabilities of GLAST to provide information on particle dark matter properties such as the mass and the dominant annihilation mode, both in virtually ``background-free'' setups and in the complex gamma-ray environment of the Galactic center; \\item study the theoretical bias that background modeling and theoretical priors (such as the dominant annihilation mode) produce in the estimation of the fundamental particle properties of dark matter from gamma-ray data. \\end{enumerate} The paper is organized as follows: sec.~\\ref{sec:dmfit} introduces the DMFIT tool and \\ref{sec:glastsim} provides details on the GLAST simulations; sec.~\\ref{sec:sources} discusses the gamma-ray sources included in the simulations, and \\ref{sec:dm} describes the particle dark matter models; sec.~\\ref{sec:dmfitdmonlly} shows examples of applications of DMFIT to gamma-ray spectra produced by dark matter annihilation only. Finally, sec.~\\ref{sec:gc} addresses the Galactic center region: we discuss there the optimal energy and angular regions for GLAST observations and the performance of GLAST at inferring particle dark matter properties from the gamma-ray spectrum from the center of the Galaxy. Our summary and conclusions are given in sec.~\\ref{sec:concl}. ", "conclusions": "\\label{sec:concl} With the detailed analysis of a specific example, the Galactic center region, this study reaffirms and reinforces the point that GLAST has the potential to play a pioneering role in the race towards the discovery of the fundamental nature of dark matter. While eagerly awaiting real data to go beyond simulations, the present theoretical study highlighted that: \\begin{itemize} \\item the nature of the EGRET source in the Galactic center, and its possible association to dark matter annihilation, will be conclusively probed by GLAST with less than one year of data; \\item the determination of the dark matter particle properties such as the mass and the dominant and sub-dominant annihilation modes is possible to a varying degree of accuracy, depending on the brightness of the source and on the knowledge of the background spectrum; \\item for realistic particle dark matter models, GLAST has the potential to (i) firmly establish the presence of a dark matter source in the Galactic center, even when it is not the brightest source in the region, (ii) estimate the dark matter particle mass to better than 10\\%, and (iii) to disentangle the occurrence of more than one annihilation mode to 99\\% confidence level; \\item if the EGRET source at the Galactic center is associated to dark matter annihilation, we showed that the optimal energy range includes all photons above 0.1 GeV within an angular region of $0.5^\\circ$; otherwise, the optimal energy range for dark matter searches is above 1 GeV, and the optimal angular region goes from $1^\\circ$ for steep dark matter profiles, to $\\sim10^\\circ$ for shallower profiles; \\item the finite spatial extent of the dark matter source at the Galactic center amounts to a systematic effect in the estimate of the normalization of the dark matter signal by a factor $\\sim 1.6$ compared to the point source approximation, even for very steep profiles; \\item the estimate of the dark matter particle properties is affected by two sources of bias: (i) the background model and (ii) assumptions on the dark matter annihilation mode; specifically, we showed that different backgrounds can lead to both under- and over-estimates of the dark matter mass, and that the inclusion of annihilation modes with hard spectra biases estimates of the mass towards lower values. \\end{itemize} We introduced and showed applications of DMFIT, a numerical package that, interfaced with any spectral fitting routine, allows one to fit gamma-ray spectra to the emission of generic WIMP models. DMFIT includes the $e^+e^-$ annihilation mode, relevant e.g. for Kaluza-Klein dark matter, and is able to fit for low neutralino masses. In addition, we provided an overview and a template of all known gamma-ray sources relevant for dark matter searches in the Galactic center region. While we do not claim here that the Galactic center is the most promising site to search for a signal from dark matter, this study indicates that GLAST data from that region can potentially be of tremendous relevance in the quest for dark matter." }, "0808/0808.0472_arXiv.txt": { "abstract": "The first phase of stellar evolution in the history of the universe may be Dark Stars, powered by dark matter heating rather than by fusion. Weakly interacting massive particles, which are their own antiparticles, can annihilate and provide an important heat source for the first stars in the the universe. This talk presents the story of these Dark Stars. We make predictions that the first stars are very massive ($\\sim 800 M_\\odot$), cool (6000 K), bright ($\\sim 10^6 L_\\odot$), long-lived ($\\sim 10^6$ years), and probable precursors to (otherwise unexplained) supermassive black holes. Later, once the initial DM fuel runs out and fusion sets in, DM annihilation can predominate again if the scattering cross section is strong enough, so that a Dark Star is born again. ", "introduction": "In October 2006, as guests of the Galileo Galilei Institute in Florence, two of us began a new line of research: the effect of Dark Matter particles on the very first stars to form in the universe. We found a new phase of stellar evolution: the first stars to form in the universe may be ``Dark Stars:'' dark matter powered rather than fusion powered. We first reported on this work in a paper (\\cite[Spolyar, Freese, \\& Gondolo 2008]{SpolyarFreeseGondolo08}) submitted to the arxiv in April 2007 (hereafter Paper I). When we presented this work at The First Stars conference in Santa Fe soon after (\\cite[Freese,Gondolo \\& Spolyar 2008]{FreeseGondoloSpolyar08}), many questions were raised, which we have addressed in the subsequent year. In this talk I review the basic ideas as well as report on the followup work we performed over the past year. The Dark Matter particles considered here are Weakly Interacting Massive Particles (WIMPs) (such as the Lightest Supersymmetric Particle), which are one of the major motivations for building the Large Hadron Collider at CERN that will begin taking data very soon. These particles are their own antiparticles; they annihilate among themselves in the early universe, leaving the correct relic density today to explain the dark matter in the universe. These particles will similarly annihilate wherever the DM density is high. The first stars are particularly good sites for annihilation because they form at high redshifts (density scales as $(1+z)^3$) and in the high density centers of DM haloes. The first stars form at redshifts $z \\sim 10-50$ in dark matter (DM) haloes of $10^6 M_\\odot$ (for reviews see e.g. \\cite[Ripamonti \\& Abel 2005]{RipamontiAbel05}, \\cite[Barkana \\& Loeb 2001]{BarkanaLoeb01}, \\cite[Bromm \\& Larson 2003]{BrommLarson03}; see also \\cite[Yoshida et al. 2006]{Yoshida_etal06}.) One star is thought to form inside one such DM halo. As our canonical values, we will use the standard annihilation cross section, $\\langle \\sigma v \\rangle = 3 \\times 10^{-26} {\\rm cm^3/sec}$, and $m_\\chi = 100$ GeV for the particle mass; but we will also consider a broader range of masses and cross-sections. Paper I found that DM annihilation provides a powerful heat source in the first stars, a source so intense that its heating overwhelms all cooling mechanisms; subsequent work has found that the heating dominates over fusion as well once it becomes important at later stages (see below). Paper I (\\cite[Spolyar, Freese, \\& Gondolo 2008]{SpolyarFreeseGondolo08}) suggested that the very first stellar objects might be ``Dark Stars,'' a new phase of stellar evolution in which the DM -- while only a negligible fraction of the star's mass -- provides the power source for the star through DM annihilation. \\begin{figure}[t] \\centerline{\\includegraphics[width=0.5\\textwidth]{sellfig.pdf}} \\caption{Adiabatically contracted DM profiles in the first protostars for an initial NFW profile (dashed line) using (a) the Blumenthal method (dotted lines) and (b) an exact calculation using Young's method (solid lines), for $M_{\\rm vir}=5 \\times 10^7 M_\\odot$, $c=2$, and $z=19$. The four sets of curves correspond to a baryonic core density of $10^4, 10^8, 10^{13},$ and $10^{16}{\\rm cm}^{-3}$. The two different approaches to obtaining the DM densities find values that differ by less than a factor of two. } \\end{figure} ", "conclusions": "The line of research we began in Florence almost two years ago is reaching a very fruitful stage of development. Dark matter can play a crucial role in the first stars. The first stars to form in the universe may be Dark Stars: powered by DM heating rather than by fusion. Our work indicates that they may be very large ($850 M_\\odot$ for 100 GeV mass WIMPs). The connections between particle physics and astrophysics are ever growing!" }, "0808/0808.2194_arXiv.txt": { "abstract": "We present the results of extensive multi-waveband monitoring of the blazar 3C~279 between 1996 and 2007 at X-ray energies (2-10 keV), optical R band, and 14.5 GHz, as well as imaging with the Very Long Baseline Array (VLBA) at 43 GHz. In all bands the power spectral density corresponds to ``red noise'' that can be fit by a single power law over the sampled time scales. Variations in flux at all three wavebands are significantly correlated. The time delay between high and low frequency bands changes substantially on time scales of years. A major multi-frequency flare in 2001 coincided with a swing of the jet toward a more southerly direction, and in general the X-ray flux is modulated by changes in the position angle of the jet near the core. The flux density in the core at 43 GHz---increases in which indicate the appearance of new superluminal knots---is significantly correlated with the X-ray flux. We decompose the X-ray and optical light curves into individual flares, finding that X-ray leads optical variations (XO) in 6 flares, the reverse occurs in 3 flares (OX), and there is essentially zero lag in 4 flares. Upon comparing theoretical expectations with the data, we conclude that (1) XO flares can be explained by gradual acceleration of radiating electrons to the highest energies; (2) OX flares can result from either light-travel delays of the seed photons (synchrotron self-Compton scattering) or gradients in maximum electron energy behind shock fronts; and (3) events with similar X-ray and optical radiative energy output originate well upstream of the 43 GHz core, while those in which the optical radiative output dominates occur at or downstream of the core. ", "introduction": "Blazars form a subclass of active galactic nuclei (AGN) characterized by violent variability on time scales from hours to years across the electromagnetic spectrum. It is commonly thought that at radio, infrared, and optical frequencies the variable emission of blazars originates in relativistic jets and is synchrotron in nature \\citep{imp88,mar98}. The X-ray emission is consistent with inverse Compton (IC) scattering of these synchrotron photons, although seed photons from outside the jet cannot be excluded \\citep{mau96,rom97,cop99,bla00,sik01,chi02,arb05}. The models may be distinguished by measuring time lags between the seed-photon and Compton-scattered flux variations. Comparison of the amplitudes and times of peak flux of flares at different wavebands is another important diagnostic. For this reason, long-term multi-frequency monitoring programs are crucially important for establishing a detailed model of blazar activity and for constraining the physics of relativistic jets. Here we report on such a program that has followed the variations in emission of the blazar 3C~279 with closely-spaced observations over a time span of $\\sim$10 years. The quasar 3C~279 \\citep[redshift=0.538;][]{bur65} is one of the most prominent blazars owing to its high optical polarization and variability of flux across the electromagnetic spectrum. Very long baseline interferometry (VLBI) reveals a one-sided radio jet featuring bright knots (components) that are ``ejected'' from a bright, presumably stationary ``core'' \\citep{jor05}. The measured apparent speeds of the knots observed in the past range from $4c$ to $16c$ \\citep{jor04}, superluminal motion that results from relativistic bulk velocities and a small angle between the jet axis and line of sight. Relativistic Doppler boosting of the radiation increases the apparent luminosity to $\\sim 10^4$ times the value in the rest frame of the emitting plasma. Characterization of the light curves of 3C~279 includes the power spectral density (PSD) of the variability at all different wavebands.The PSD corresponds to the power in the variability of emission as a function of time scale. \\citet{law87} and \\citet{mch87} have found that the PSDs of many Seyfert galaxies, in which the continuum emission comes mainly from the central engine, are simple ``red noise'' power laws, with slopes between $-1$ and $-2$. More recent studies indicate that some Seyfert galaxies have X-ray PSDs that are fit better by a broken power law, with a steeper slope above the break frequency \\citep{utt02, mch04, mar03, ede99, pou01}. This property of Seyferts is similar to that of Galactic black hole X-ray binaries (BHXRB), whose X-ray PSDs contain one or more breaks \\citep{bel90,now99}. However, in blazars most of the X-rays are likely produced in the jets rather than near the central engine as in BHXRBs and Seyferts. It is unclear {\\it a priori} what the shape of the PSD of nonthermal emission from the jet should be, a question that we answer with the dataset presented here. The raw PSD calculated from a light curve combines two aspects of the dataset: (1) the intrinsic variation of the object and (2) the effects of the temporal sampling pattern of the observations. In order to remove the latter, we apply a Monte-Carlo type algorithm based on the ``Power Spectrum Response Method'' (PSRESP) of \\citet{utt02} to determine the intrinsic PSD (and its associated uncertainties) of the light curve of 3C~279 at each of three wavebands. Similar complications affect the determination of correlations and time lags of variable emission at different wavebands. Uneven sampling, as invariably occurs, can cause the correlation coefficients to be artificially low. In addition, the time lags can vary across the years owing to physical changes in the source. In light of these issues, we use simulated light curves, based on the underlying PSD, to estimate the significance of the derived correlation coefficients. In {\\S}2 we present the observations and data reduction procedures, while in {\\S}3 we describe the power spectral analysis and its results. In {\\S}4 we cross-correlate the light curves to determine the relationship of the emission at different wavebands. Finally, in {\\S}5 we discuss and interpret the results, focusing on the implications regarding the location of the nonthermal radiation at different frequencies, as well as the physical processes in the jet that govern the development of flux outbursts in blazars. ", "conclusions": "This paper presents well-sampled, decade-long light curves of 3C~279 between 1996 and 2007 at X-ray, optical, and radio wavebands, as well as monthly images obtained with the VLBA at 43 GHz. We have applied an algorithm based on a method by \\citet{utt02} to obtain the broadband PSD of nonthermal radiation from the jet of 3C~279. Cross-correlation of the light curves allows us to infer the relationship of the emission across different wavebands, and we have determined the significance of the correlations with simulated light curves based on the PSDs. Analysis of the VLBA data yields the times of superluminal ejections and reveals time variations in the position angle of the jet near the core. We have identified 13 associated pairs of X-ray and optical flares by decomposing the light curves into individual flares. Comparison of the observed radiative energy output of contemporaneous X-ray and optical flares with theoretical expectations has provided a quantitative evaluation of synchrotron and SSC models. We have discussed the results by focusing on the implications regarding the location of the nonthermal radiation at different frequencies, physical processes in the jet, and the development of disturbances that cause outbursts of flux density in blazars. Our main conclusions are as follows: \\\\ (1) The X-ray, optical, and radio PSDs of 3C~279 are of red noise nature, i.e., there is higher amplitude variability at longer time scales than at shorter time scales. The PSDs can be described as power laws with no significant break, although a break in the X-ray PSD at a variational frequency $\\lesssim 10^{-8}$ Hz cannot be excluded at a 95\\% confidence level.\\\\ (2) X-ray variations correlate with those at optical and radio wavebands, as expected if nearly all of the X-rays are produced in the jet. The X-ray flux correlates with the projected jet direction, as expected if Doppler beaming modulates the mean X-ray flux level.\\\\ (3) X-ray flares are associated with superluminal knots, with the times of the latter indicated by increases in the flux of the core region in the 43 GHz VLBA images. The correlation has a broad peak at a time lag of $130^{+70}_{-45}$ days, with X-ray variations leading.\\\\ (4) Analysis of the X-ray and optical light curves and their interconnection indicates that the X-ray flares are produced by SSC scattering and the optical flares by the synchrotron process. Cases of X-ray leading the optical peaks can be explained by an increase in the time required to accelerate electrons to the high energies needed for optical synchrotron emission. Time lags in the opposite sense can result from either light-travel delays of the SSC seed photons or gradients in maximum electron energy behind the shock fronts. \\\\ (5) The switch to optical-leading flares during the major multi-frequency outburst of 2001 coincided with a decrease in the apparent speeds of knots from 16-17$c$ to 4-7$c$ and a swing toward the south of the projected direction of the jet near the core. This behavior, as well as the high amplitude of the outburst, can be explained if the redirection of the jet (only a 1$\\degr$-2$\\degr$ change is needed) caused it to point closer to the line of sight than was the case before and after the 2001-02 outburst. \\\\ (6) Contemporaneous X-ray and optical flares with similar radiative energy output originate closer to the base of the jet, where the cross-section of the jet is smaller, than do flares in which the optical energy output dominates. This is supported by the longer time delay in the latter case. This effect is caused by the lower electron density and magnetic field and larger cross-section of the jet as the distance from the base increases. Further progress in our understanding of the physical structures and processes in compact relativistic jets can be made by increasing the number of wavebands subject to intense monitoring. Expansion of such monitoring to a wide range of $\\gamma$-ray energies will soon be possible when GLAST scans the entire sky several times each day. When combined with similar data at lower frequencies as well as VLBI imaging, more stringent tests on models for the nonthermal emission from blazars will be possible." }, "0808/0808.2477_arXiv.txt": { "abstract": "It has been argued that low-luminosity dwarf galaxies are the dominant source of ionizing radiation during cosmological reionization. The fraction of ionizing radiation that escapes into the intergalactic medium from dwarf galaxies with masses less than $\\sim$$10^{9.5}$ solar masses plays a critical role during this epoch. Using an extensive suite of very high resolution (0.1 pc), adaptive mesh refinement, radiation hydrodynamical simulations of idealized and cosmological dwarf galaxies, we characterize the behavior of the escape fraction in galaxies between $3 \\times 10^6$ and $3 \\times 10^9$ solar masses with different spin parameters, amounts of turbulence, and baryon mass fractions. For a given halo mass, escape fractions can vary up to a factor of two, depending on the initial setup of the idealized halo. In a cosmological setting, we find that the time-averaged photon escape fraction always exceeds 25\\% and reaches up to 80\\% in halos with masses above $10^8$ solar masses with a top-heavy IMF. The instantaneous escape fraction can vary up to an order of magnitude in a few million years and tend to be positively correlated with star formation rate. We find that the mean of the star formation efficiency times ionizing photon escape fraction, averaged over all atomic cooling ($T_{\\rm vir} \\ge 8000~$K) galaxies, ranges from 0.02 for a normal IMF to 0.03 for a top-heavy IMF, whereas smaller, molecular cooling galaxies in minihalos do not make a significant contribution to reionizing the universe due to a much lower star formation efficiency. These results provide the physical basis for cosmological reionization by stellar sources, predominately atomic cooling dwarf galaxies. ", "introduction": "In most calculations of cosmological reionization, it has to be assumed that the product of star formation efficiency and hydrogen ionizing photon escape fraction is $\\ge 0.01$ \\citep[e.g.][]{Gnedin00, Cen03a, Cen03b, Wyithe03a, Wyithe03b, Venkatesan03, Somerville03, Chiu03, Haiman03, Ciardi03, Sokasian03, Sokasian04, Wyithe07, Srbinovsky08} in order to reionize the universe early enough to be consistent with the Wilkinson Microwave Anisotropy Probe (\\textit{WMAP}) observations \\citep{Spergel07, Komatsu08}, if stars produce the majority of ionizing photons. \\citet{Gnedin00} suggested that, based on Local Group dwarf galaxies, the star formation efficiency at high redshift is $\\sim 4\\%$, consistent with theoretical works \\citep[e.g.][]{Krumholz05, Krumholz07}. Thus, unless star formation efficiency is unusually high (i.e., $\\ge 10\\%$) at high redshift, this requirement seems to suggest a high ionizing photon escape fraction from high redshift galaxies, $f_{\\rm esc} \\ge 10\\%$, may be necessary in the context of stellar reionization within the standard cold dark matter model. However, $f_{\\rm esc} \\ge 10\\%$ is by no means the norm, at least from observations at lower redshifts. Approximately 6\\% of ionizing radiation escape from the Milky Way \\citep{Bland99, Putman03}. For local starburst galaxies \\citet{Hurwitz97} gave $f_{\\rm esc}\\le 0.032,0.052, 0.11$ ($2\\sigma$) for Mrk 496, Mrk 1267, and IRAS 08339+6517 ($\\le 0.57$ in the case of Mrk 66). \\citet{Deharveng01} gave an escape fraction of $f_{\\rm esc}<0.062$ for Mrk 54. \\citet{Heckman01} found $f_{\\rm esc} \\le 0.06$. \\citet{Bergvall06} found $f_{\\rm esc}\\sim 0.04-0.1$ for a local extreme starburst dwarf, the Blue Compact Galaxy Haro 11. \\citet{Chen07} placed a 95\\% upper limit of $f_{\\rm esc}\\le 0.075$ for star forming regions hosting gamma-ray bursts at $z\\ge 2$. For Lyman break galaxies at $z\\sim 3$, \\citet{Shapley06} found $f_{\\rm esc}\\sim 0.03$. \\citet{Siana07} examined starburst galaxies at $z \\sim 1.3$ and found that less than 20\\% have relative escape fractions% \\footnote{This quantity is the absolute escape fraction that is corrected for dust attenuation at 1500 \\AA~and is a common measure in observations of escaping Lyman continuum.} near unity and a global absolute escape fraction of $\\lsim 0.04$. However in this study, some galaxies have upper limits of $f_{\\rm esc, rel}$ as low as 0.08. \\citet{Inoue06} recently compiled available observations over a wide range of redshift and come to the conclusion that there is a trend of increasing $f_{\\rm esc}$ with redshift from $\\le 0.01$ at $z\\le 1$ to $\\sim 0.1$ at $z\\ge 4$. Some theoretical works also seem to point to low values for $f_{\\rm esc}$. Theoretical models of our own Galaxy with realistic star formation mode by \\citet{Dove00} give an estimate of $f_{\\rm esc}\\sim 0.06$ or lower. \\citet{Ciardi02} studied the effect of gas inhomogeneities on $f_{\\rm esc}$ for Milky Way like galaxies using simulations with radiative transfer and concluded that $f_{\\rm esc}$ depends on the density structure as well as star formation rate (SFR). In addition, porosity in the interstellar medium (ISM) that is caused by SN mechanical feedback may provide additional channels in which radiation could escape, thus increasing the UV escape fraction \\citep{Clarke02}. \\citet{Wood00} argued that $f_{\\rm esc}\\le 0.01$ for galaxies at $z\\sim 10$. \\citet{Ricotti00} found that $f_{\\rm esc}\\ge 0.1$ only for halos of total mass less than $\\sim 10^7\\msun$ at $z\\ge 6$ with $f_{\\rm esc}$ dropping precipitously for larger halos. \\citet{Fujita03}, using ZEUS-3D simulations, found that $f_{\\rm esc}\\le 0.1$ from disks of dwarf starburst galaxies of total mass $M\\ge 10^{8-10}\\msun$. \\citet[][hereafter RS06 and RS07]{Razoumov06, Razoumov07} concluded that $f_{\\rm esc}$ = 0.01--0.1 in several young galaxies with $\\mvir = 10^{12-13} \\Ms$ at $z \\sim 3$ in their simulations with radiative transfer using adaptive ray tracing, agreeing with the observations presented in \\citet{Inoue06}. \\citet[][hereafter GKC08]{Gnedin08}, using detailed hydrodynamic simulations with radiative transfer, found $f_{\\rm esc}\\sim 0.01-0.03$ for galaxies of total mass $M\\ge 10^{11}\\msun$ at $z=3-5$. In the relevant redshift range for cosmological reionization, $z=6-15$, most of the ionizing radiation due to stars comes from dwarf galaxies of $M\\le 10^{8-9}\\msun$ in the standard cold dark matter model \\citep{Barkana01}. This purpose of this paper is to investigate how $f_{\\rm esc}$ depends on various physical parameters expected in realistic cosmological settings using detailed adaptive mesh refinement (AMR) simulations coupled with adaptive 3D ray-tracing radiative transfer, focusing on galaxies with total mass $M=10^{6.5}-10^{9.5}\\msun$ at $z\\ge 6$. We study the dependence of $f_{\\rm esc}$ on four physical parameters: mass of the galaxy, spin parameter, baryonic mass fraction and turbulent energy. Our work is an extension of GKC08 and \\citet{Fujita03}, by exploring the dependence of $f_{\\rm esc}$ on some important physical parameters aforementioned, by including lower mass galaxies and by employing very high resolution ($0.1$pc) simulations. In addition, importantly, we allow for disparate lifetimes of stars of different masses, of their subsequent explosive energy (for those stars that become supernovae) input into the ISM. We first describe our radiation hydrodynamics simulations of isolated and cosmological halos in \\S\\ref{sec:sims}. There we also describe our algorithm for star formation and feedback that considers a multi-phase ISM and resolved molecular clouds. In \\S\\ref{sec:sf} and \\S\\ref{sec:fesc}, we present the star formation rates and history and the resulting escape fraction of UV radiation, respectively. Next in \\S\\ref{sec:disc}, we further compare our results with previous work and discuss effects from any neglected physical processes in our simulations and the implications of our results on reionization scenarios. We summarize our work in the last section. ", "conclusions": "\\label{sec:disc} The fraction of ionizing radiation that escapes high-redshift galaxies is an important value to quantify in order to characterize cosmological reionization. In this section, we first discuss the differences and similarities, along with their respective causes, between our results and previous work. Next we discuss any implications and agreements with semi-analytic reionization models. We last elaborate on possible influences from neglected physical processes. \\subsection{Possible Physical and Computational Dependencies} Our results suggest that $f_{\\rm esc}$ is higher than 0.1 in such galaxies, but we must first understand why our $f_{\\rm esc}$ values differ from the more massive galaxies presented in the simulations of \\citet{Razoumov06, Razoumov07} and \\citet{Gnedin08}. Below we discuss five possible causes. \\textit{Galaxy morphology and environment}--- The galaxies simulated here exhibit an irregular morphology and did not form a rotationally-supported disk, whereas the simulations of RS06, RS07, and GKC08 all studied disk galaxies. As mentioned in \\citet{Fujita03}, a turbulent and clumpy ISM may allow for radiation to escape more easily into the IGM. Compared to a disk configuration, there are more low-density escape routes, i.e. porosity, for the radiation because of the clumpy nature of the gas \\citep{Clarke02}. Furthermore, the stellar orbits are not confined within a disk and can reach the outskirts of the galaxy at apocenter, where gas densities are much lower than the galactic center and ionizing radiation can escape into the IGM at a much greater rate. Perhaps this irregular morphology is only initially present in dwarf galaxies because radiative feedback shifts the angular momentum distribution to higher values \\citep{Wise08a}. A fraction of this material will return to the galaxy and aid in disk formation in higher mass galaxies, such as the ones studied in RS06, RS07, and GKC08. Furthermore in rare halos, the intersecting filaments tend to be dense and thin, where a halo with equal mass at lower redshift would be contained in a large filament \\citep{Ocvirk08}. This redshift dependency on halo environment could affect \\fesc~at $r > r_{\\rm vir}$, but should not be apparent at \\rvir. \\textit{Galactic mass}--- GKC08 find that $f_{\\rm esc}$ increases as much as 3 orders of magnitude from a halo mass of \\tento{10}~to \\tento{11}\\Ms. The cold \\ion{H}{1} disk in their lower mass galaxies tend to be more vertically extended, which brings about this precipitous drop. If this trend continues to lower masses, the values of $f_{\\rm esc}$ would be negligible in our galaxies, but we see no such continuation. Our results suggest that the radiative feedback in dwarf galaxies with $V_c \\lsim 35 \\kms$ have a profound impact on the ISM by driving outflows and preventing any disk formation, increasing \\fesc. Perhaps this circular velocity is a critical turning point below which $f_{\\rm esc}$ is large ($>0.1$) because of dynamical effects from radiative feedback. In more massive galaxies, SN feedback provides the main impetus for driving outflows. \\textit{Star formation rates}--- One can argue that our simulations only capture the initial collapse and an artificially strong starburst in high-redshift galaxies; however, after $\\sim$20 Myr of star formation, this collapse is reversed by stellar feedback and SFRs stabilize. In addition our most massive galaxy ($\\mvir = 4 \\times 10^9 \\Ms$) has an average SFR of 0.2 \\Ms~yr$^{-1}$, which is comparable to the SFRs found in the $\\mvir \\sim 10^{10} \\Ms$ galaxies in GKC08. The star formation efficiencies ($M_\\star/M_{\\rm{gas}}$) of our galaxies range from 5--10\\%, also agreeing with previous galaxy formation simulations. Hence we do not think our treatment of star formation affects the SFRs and resulting UV escape fraction. \\textit{Stellar IMF}--- We experimented with both normal ($N_\\gamma = 2,600$) and top-heavy ($N_\\gamma = 26,000$) IMFs and found that a normal IMF causes $f_{\\rm esc}$ values to be lower by 10--75\\%. The differences in how the gas and ensuing star formation reacts to radiative feedback most likely causes this spread. The normal IMF is similar to the ones used in all previously quoted theoretical works, and thus we do not expect our choice in IMF to cause our high UV escape fractions. One minor shortcoming of our simulations is that we keep the SN feedback strength equal in our normal and top-heavy IMFs. Although this helps us localize the cause of any changes to a different specific luminosity, the strength of SN feedback should also affect the gas distribution, SFR, and \\fesc. \\textit{Resolution}--- We can afford a spatial resolution of $\\sim$0.1 physical pc in all of our simulations. However in larger galaxies, it becomes increasingly difficult to maintain this resolution, especially in cosmological simulations. The highest resolution simulation in RS06 and RS07 use a gravitational softening length of 330 $h^{-1}$ pc, and their radiative transfer grid is refined so that 10 particles occupy each cell. GKC08 use AMR simulations with a maximum resolution of 50 physical pc. \\citet{Fujita03} have a fixed resolution of 0.28 pc in their $z = 8, M_d = 10^8 \\Ms$ simulation, similar to our resolution. The $f_{\\rm esc}$ values of our simulations and \\citeauthor{Fujita03} are in agreement. Regular and giant molecular clouds have radii of 2--20 and 10--60 pc \\citep[see][for a review]{MacLow04}, respectively, and are usually accounted with subgrid physics in cosmological simulations. However our simulations capture the turbulent and clumpy nature of the ISM, while resolving such molecular clouds. As discussed before, the turbulent ISM and its porosity may play a key role in allowing radiation to escape \\citep{Clarke02}. In addition, high-resolution simulations with radiative transfer can model the fragmentation of ionization fronts \\citep[e.g.][]{Fujita03, Whalen08a}, which can further assist in boosting UV escape fractions. \\subsection{Impact on reionization models} \\label{sec:models} In reionization models, whether it be based on Press-Schetcher formalism or N-body simulations, the product $\\fescs \\equiv f_\\star \\times \\fesc$ ultimately dictates the absolute number of photons that escape from each halo into the IGM. When calibrating these models against \\textit{WMAP} observations \\citep{Spergel07, Komatsu08} and \\lya~transmission from $z \\sim 6$ quasars \\citep[e.g.][]{Bolton07, Srbinovsky08}, the values of $\\ge 0.01$ are usually required. Recently, \\citeauthor{Srbinovsky08} found that \\fesc~must lie within the range 0.1--0.25 with a best-fit of $f_\\star = 0.11$ if all halos with masses $\\gsim 10^9 \\Ms$ contribute to reionization. Furthermore, they conclude that even the smallest atomic line cooling halos ($\\mvir \\gsim 10^8 \\Ms$) have $\\fesc \\sim 0.05$. Our results support this scenario where low-luminosity galaxies are the main contributors to reionization. We plot \\fescs~from idealized and cosmological halos in Figure~\\ref{fig:fesc_cstar}. We calculate $f_\\star$ using the initial gas mass in halo at the initial time; thus we are overestimating $f_\\star$ up to a factor of $\\sim$1.5 because the halos experience mass accretion in the 100 Myr that we have simulated. Recall that none of the halos presented here undergo a major merger. In our least massive cosmological halos with $\\mvir \\le 3 \\times 10^7 \\Ms$, \\fescs~increases rapidly with respect to halo mass from \\tento{-3}~to \\tento{-2} because atomic hydrogen line cooling becomes efficient in these halos. \\fescs~is also an order of magnitude higher than the idealized cases. As discussed in \\S\\ref{sec:sfr}, the additional gas accretion from filaments results in higher SFRs or equivalently $f_\\star$. In atomic cooling halos ($\\tvir > 8000$ K), we find \\fescs~to always be $\\ge 0.01$, sufficiently high enough to meet the ``critical'' value to match reionization constraints, with an average of 0.033 and 0.026 for top-heavy and normal IMFs, respectively. When integrated over a luminosity function with a faint-end slope of --1.7 \\citep[e.g.][]{Bouwens07}, the average of \\fescs~in atomic cooling halos is 0.029 and 0.021 for top-heavy and normal IMFs, respectively. If we include all halos, this average drops to $4.1 \\times 10^{-3}$ because the low-mass halos are more abundant and cannot cool and form stars as efficiently. In Figure~\\ref{fig:fesc_cstar_trends}, we show the behavior of \\fescs~with halo mass in idealized halos grouped by halo parameters, similar to Figure~\\ref{fig:trends}. There is no apparent trends with respect to turbulent energy; however, halos with higher spin parameters result in smaller values of \\fescs~in high mass halos that form a rotationally supported disk. The halos with $f_b = 0.1$ have the highest values of \\fescs, and gas-poor halos are consistently lower in all halo masses. \\begin{figure}[t] \\begin{center} \\epsscale{1.15} \\plotone{f11_color} \\caption{\\label{fig:fesc_cstar} The product of stellar mass fraction ($M_{\\rm star}/M_{\\rm gas}$) and escape fraction of ionizing photons ($f_{\\rm esc}$) from idealized halos and cosmological halos with a top-heavy IMF (filled diamonds) and normal IMF (open diamonds) as a function of halo mass. Idealized halos with a normal IMF are plotted as large circles. This product is a key quantity in semi-analytical reionization models in determining the amount of escaping radiation per collapsed gas fraction.} \\end{center} \\end{figure} \\begin{figure*}[t] \\begin{center} \\epsscale{1.0} \\plotone{f12_color} \\caption{\\label{fig:fesc_cstar_trends} Relative changes in the product of stellar mass fraction ($M_{\\rm star}/M_{\\rm gas}$) and escape fraction of ionizing photons ($f_{\\rm esc}$) from all idealized models grouped by initial halo parameters: turbulent energy (\\textit{left}), spin parameter (\\textit{middle}), and baryon mass fraction (\\textit{right}) when compared to the ``control halos'' with ($f_{\\rm turb}, \\lambda, f_b$) = (0.25, 0.04, 0.1).} \\end{center} \\end{figure*} \\subsection{Effects from the Ultraviolet Background} Our simulations accurately track the evolution of the \\ion{H}{2} regions and how they breakout into the IGM. Here we discuss an important process that was neglected in our simulations, external feedback from the UV background (UVB). Our models sample halos that are primarily dependent on \\hh~cooling ($\\tvir \\lsim 8000$ K) and atomic line cooling in more massive halos. Gas condensation in the lower mass \\hh~cooling halos can be delayed by a \\hh~dissociating UV (Lyman-Werner) background between 11.2 and 13.6 eV \\citep{Machacek01, Wise07b, OShea08}. The Lyman-Werner background thus increases cooling times in the centers of such halos. As a result, the minimum mass of a star-forming halo increases with the Lyman-Werner background intensity. This will not affect the global halo properties but may suppress the SFR in these halos. The Lyman-Werner background becomes less of an issue in atomic line cooling halos as \\lya~cooling provides ample amounts of free electrons for \\hh~cooling, and they become self-shielding to this radiation \\citep{Wise08a}. We now consider any photo-heating from a hydrogen ionizing UVB, which can partially suppress gas accretion and thus star formation in dwarf galaxies \\citep[e.g.][]{Efstathiou92, Shapiro94, Thoul96}. Here the Jeans ``filtering mass'' \\citep{Gnedin98} accurately describes the minimum halo mass that can cool and collapse given an IGM thermal history \\citep{Gnedin00, Wise08b}. The Jeans filtering mass is calculated by analyzing the linear evolution of overdensities in the presence of Jeans smoothing. Because the low-density IGM has a dynamical time on the order of a Hubble time, it slowly reacts to any photo-heating, and the filtering mass can be thought of some running time-average of the Jeans mass. Thus at some redshift, the Jeans and filtering mass may differ substantially. \\citet{Thoul96} originally showed photo-heating could totally suppress any star formation in low-redshift dwarf galaxies with $V_c \\lsim 35 \\kms$ and somewhat lower SFRs in galaxies up to 100 \\kms. At $z \\gsim 3$, halos are less susceptible to negative feedback from photo-heating because of their higher average densities and thus cooling rates \\citep{Dijkstra04}. Closer inspection of collapsing halos in radiation hydrodynamics simulations also shows high-redshift, low-mass halos can collapse and is indeed regulated by the filtering mass. Prior to reionization, \\citet{Gnedin00} showed that the filtering mass smoothly increases from $\\sim$\\tento{6}~\\Ms~before any star formation occurs to \\tento{9}\\Ms~at $z = 6$. \\citet{Wise08b} calculated the filtering mass in regions that were ionized by Population III stars and found that it gradually increases the filtering mass to $\\sim$$3 \\times 10^7 \\Ms$ at $z \\sim 15$ around biased regions. Recently, \\citet{Mesigner08} studied how photo-heating from an inhomogeneous UVB suppresses gas cooling and star formation. They found that at $z = 10$ an UVB intensity of $J_{21} \\sim 0.1$ is necessary to completely suppress gas collapse in $\\mvir = 10^8 \\Ms$ halos, where $J_{21}$ is in units of \\tento{-21}~\\emis. In their semi-numerical simulations at $z = 10$, they showed this intensity occurs in a volume fraction of $<1\\%$, suggesting that star formation is widespread in these low-mass halos even in the advanced stages of reionization. They also illustrate how the collapsed gas fraction steadily decreases with increasing UVB. Even at $z = 7$ in a \\tento{8}\\Ms~halo, approximately 40\\% of the gas is retained when compared to the no feedback case. This gas mass loss will directly affect star formation and perhaps the subsequent UV escape fraction. Using the results of \\citeauthor{Mesigner08} and the dependence of \\fescs~on gas fraction (see Fig. \\ref{fig:fesc_cstar_trends}), our results from cosmological halos can be adjusted to account for external feedback from the UVB. For example, at $z = 10$ with their fiducial model, their semi-numerical simulations show that $J_{21} \\sim 0.01$ is most common, which decreases the gas fraction of \\tento{8}, \\tento{8.5}, \\tento{9}\\Ms~halos to 0.2, 0.6, 0.85, respectively, of the gas fraction without any UVB. From Fig. \\ref{fig:fesc_cstar_trends}, we can see that \\fescs~drops by $\\sim$50\\% when the gas fraction is $\\le$0.075 in idealized halos. We would expect a similar decrease in halos that lose gas from UVB feedback, resulting in $\\fescs \\sim 0.02$ for halos in this mass range. \\subsection{Prior Star Formation Episodes} We assumed that the halos were unaffected by any prior stellar feedback, but in this paper, we have also shown that low-mass halos are susceptible to gas ``blow-out'' caused by radiative feedback. This effect should be evident starting with Population III star formation \\citep[e.g.][]{Yoshida07, Wise08a} in halos with circular velocities up to $V_c \\sim 35 \\kms$. For example, \\citeauthor{Wise08a} found that gas fractions of dwarf galaxies with $\\mvir \\sim 3 \\times 10^7 \\Ms$ at redshift 15 have been decreased to $\\le0.1$. The cosmological halos presented here, which have $f_b \\sim 0.13$, have $f_{\\rm esc} \\sim 0.5$ with a top-heavy IMF; however if $f_b$ were lower from previous gas ``blow-out'', the trends shown in Figure \\ref{fig:trends} suggest that the escape fraction in these halos should be even higher, but their values of \\fescs~(Fig. \\ref{fig:fesc_cstar_trends}) will decrease because of lower SFRs. To further understand and quantify the consequences of these neglected processes, semi-analytical reionization models, semi-numerical or N-body + radiative transfer simulations are needed, as discussed in \\S\\ref{sec:models}. In future work, we plan to study this in detail by following the hierarchical assembly of high-redshift dwarf galaxies in cosmological simulations, including both metal-free and metal-enriched star formation and feedback. We presented results from an extensive suite of very high resolution ($0.1~$pc) AMR radiation hydrodynamics simulations that focus on the UV escape fraction from isolated halos and cosmological halos. The latter cases were extracted from two large-scale cosmological simulations. These simulations accurately track the evolution of \\ion{H}{2} regions while including radiative and SNe feedback on the surrounding medium. The halo masses studied in this work span from $3 \\times 10^6$ to $3 \\times 10^9 \\Ms$. We used the isolated halos to gauge if the escape fraction depends on the turbulent fractional energy, spin parameter, and baryon mass fraction of the halo. The cosmological halos provide a good estimate of realistic escape fractions of high-redshift dwarf galaxies. We also investigated how a top-heavy IMF and a normal IMF affects the escape fraction. Our main findings in this paper are \\medskip 1. Radiative feedback from massive stars, primarily arising from D-type fronts, is most effective at ejecting gas from halos with masses less than $10^{8.5} \\Ms$ ($V_c = 35 \\kms$). 2. Radiation preferentially escapes through channels with low column densities, creating anisotropic \\ion{H}{2} and a highly varying $f_{\\rm esc}$ along different lines of sight, agreeing with previous work. 3. For a given halo mass, escape fractions in isolated halos have a standard deviation of 0.2, arising from differing physical halo parameters. The largest effect comes from the baryon mass fraction, where $f_{\\rm esc}$ is lower in gas-rich halos with masses below \\tento{8} \\Ms. In gas-rich halos with masses larger than \\tento{8} \\Ms, star formation is more intense, overcoming any large neutral fraction it must ionize and resulting in higher values of $f_{\\rm esc}$ than halos with smaller gas fractions. 4. High-redshift dwarf galaxies with $\\mvir > 10^7 \\Ms$, a top-heavy IMF, and irregular morphology have $0.25 \\le f_{\\rm esc} \\le 0.8$, which we determined from our simulations of cosmological halos. A normal IMF decreases $f_{\\rm esc}$ to 0.05--0.1 in halos with $\\mvir < 10^{7.5} \\Ms$ and $f_{\\rm esc} \\sim 0.4$ in more massive halos, which have maximum SFRs spanning almost 2 orders of magnitude. 5. Escape fractions are dependent not only on the current SFR but on the photo-heating and dispersion of gas, following feedback from previous episodes of star formation. Values of $f_{\\rm esc}$ can vary up to an order of magnitude in a few million years, i.e. the dynamical time of a molecular cloud on which variations in SFRs can occur. 6. The mean of the product of star formation efficiency and ionizing photon escape fraction, averaged over all atomic cooling ($\\tvir \\ge 8000~$K) galaxies, ranges from $0.02$ for a normal IMF to $0.03$ for a top-heavy IMF. Smaller, molecular cooling galaxies in minihalos are not significant contributors to reionizing the universe primarily because of a much lower star formation efficiency in minihalos than in atomic cooling halos. \\medskip The high escape fraction of UV radiation has important implications on reionization, allowing a large amount of radiation from low-luminosity dwarf galaxies to freely propagate into the IGM. Although our simulations miss the entire assembly of halos and their complete star formation history, they are robust in following the dynamics of star formation and feedback during the 100 Myr studied and suggest that escape fraction in these galaxies are larger than previously assumed." }, "0808/0808.4121_arXiv.txt": { "abstract": "The recent {\\it Spitzer} detections of the 9.7$\\mum$ Si--O silicate emission in type 1 AGNs provide support for the AGN unification scheme. The properties of the silicate dust are of key importance to understanding the physical, chemical and evolutionary properties of the obscuring dusty torus around AGNs. Compared to that of the Galactic interstellar medium (ISM), the 10$\\mum$ silicate emission profile of type 1 AGNs is broadened and has a clear shift of peak position to longer wavelengths. In literature this is generally interpreted as an indication of the deviations of the silicate composition, size, and degree of crystallization of AGNs from that of the Galactic ISM. In this {\\it Letter} we show that the observed peak shift and profile broadening of the 9.7$\\mum$ silicate emission feature can be explained in terms of porous composite dust consisting of ordinary interstellar amorphous silicate, amorphous carbon and vacuum. Porous dust is naturally expected in the dense circumnuclear region around AGNs, as a consequence of grain coagulation. ", "introduction": "Dust is the cornerstone of the unification theory of active galactic nuclei (AGNs). This theory proposes that all AGNs are essentially ``born equal'' -- all types of AGNs are surrounded by an optically thick dust torus and are basically the same object but viewed from different lines of sight (see e.g. Antonucci 1993; Urry \\& Padovani 1995): type 1 AGNs, which always display broad hydrogen emission lines in the optical and have no obvious obscuring effect, are viewed face-on which allows a direct view of the central nuclei, while type 2 AGNs are viewed edge-on with most of the central engine and broad line regions being hidden by the obscuring dust. Silicate dust, a major solid species in the Galactic interstellar medium (ISM) as revealed by the strong 9.7$\\mum$ and 18$\\mum$ bands (respectively ascribed to the Si--O stretching and O--Si--O bending modes in some form of silicate material, e.g. olivine Mg$_{2x}$Fe$_{2-2x}$SiO$_4$), has also been detected in AGNs both in {\\it emission} and in {\\it absorption} (see Li 2007 for a review). The first detection of the silicate {\\it absorption} feature in AGNs was made at 9.7$\\mum$ for the prototypical Seyfert 2 galaxy NGC\\,1068 (Rieke \\& Low 1975; Kleinmann et al.\\ 1976), indicating the presence of a large column of silicate dust in the line-of-sight to the nucleus. It is known now that most of the type 2 AGNs display silicate {\\it absorption} bands (e.g. see Roche et al.\\ 1991, Siebenmorgen et al.\\ 2004, Hao et al.\\ 2007, Spoon et al.\\ 2007, Roche et al.\\ 2007) which is expected from the AGN unified theory -- for a centrally heated optically thick torus viewed edge-on, the silicate features should be in absorption. For type 1 AGNs viewed face-on, one would expect to see the silicate features in {\\it emission} since the silicate dust in the surface of the inner torus wall will be heated to temperatures of several hundred kelvin to $\\simali$1000$\\K$ by the radiation from the central engine, allowing for a direct detection of the 9.7$\\mum$ and 18$\\mum$ silicate bands emitted from this hot dust. However, their detection has only very recently been reported in a number of type 1 AGNs covering a broad luminosity range, thanks to {\\it Spitzer} (Hao et al.\\ 2005, Siebenmorgen et al.\\ 2005, Sturm et al.\\ 2005, Weedman et al.\\ 2005, Shi et al.\\ 2006, Schweitzer et al.\\ 2008). It is worth noting that the silicate emission features have recently also been detected in type 2 QSOs (Sturm et al.\\ 2006, Teplitz et al.\\ 2006). \\begin{figure} \\begin{center} \\psfig{figure=f1.cps, width=2.5in, angle=270} \\end{center} \\caption{\\label{fig:sil_obs_spec} Comparison of the silicate emission features of the quasar 3C\\,273 (Hao et al.\\ 2005) and the low-luminosity AGN NGC\\,3998 (Sturm et al.\\ 2005) with the silicate absorption feature of the ISM toward the Galactic center object Sgr A$^{\\ast}$ (Kemper et al.\\ 2004). Most notably is the long wavelength shift of the peak positions of 3C\\,273 and NGC\\,3998 in comparison with the ISM profile. All profiles are normalized to their peak values. } \\end{figure} Compared to that of the Galactic ISM, the 9.7$\\mum$ silicate emission profiles of some AGNs appear ``anomalous''. As illustrated in Figure \\ref{fig:sil_obs_spec}, both quasars (high luminosity counterparts of Seyfert 1 galaxies; Hao et al.\\ 2005, Siebenmorgen et al.\\ 2005) and the low-luminosity AGN NGC\\,3998 (Sturm et al.\\ 2005) exhibit silicate emission peaks at a much longer wavelength ($\\simali$10--11.5$\\mum$), inconsistent with the ``standard'' silicate ISM dust (which peaks at $\\simali$9.7$\\mum$). The 9.7$\\mum$ feature of NGC\\,3998 is also much broader than that of the Galactic ISM (Sturm et al.\\ 2005).\\footnote{% We should note that the width of the interstellar 9.7$\\mum$ Si--O absorption feature is not universal, but varies from one sightline to another (see Draine 2003). Generally speaking, it is relatively narrow in diffuse clouds and broad in molecular clouds (Bowey et al.\\ 1998). In contrast, its peak wavelength is relatively stable. } The deviations of the silicate emission profiles of type 1 AGNs from that of the Galactic ISM dust are generally interpreted as an indication of the differences between AGNs and the Galactic ISM in the composition, size distribution, and degree of crystallization of the silicate dust (Sturm et al.\\ 2005). However, we show in this {\\it Letter} that the observed peak shift and profile broadening of the 9.7$\\mum$ silicate emission features of AGNs can be explained in terms of porous composite dust consisting of ordinary interstellar amorphous silicate, amorphous carbon\\footnote{% There must exist a population of carbonaceous dust in the AGN torus, as revealed by the detection of the 3.4$\\mum$ absorption feature, attributed to the C--H stretching mode in saturated aliphatic hydrocarbon dust (see Li 2007 for a review). Whether the bulk form of the carbonaceous component in AGNs is amorphous carbon or hydrogenated amorphous carbon is not clear. } and vacuum, without invoking an exotic dust composition or size distribution. In the dense circumnuclear region around AGNs, a porous structure is naturally expected for the dust formed through the coagulation of small silicate and carbonaceous grains. ", "conclusions": "} It is well-known theoretically that as the size of a silicate grain increases, the 9.7$\\mum$ Si--O feature becomes wider with its peak shifted to longer wavelengths (see Fig.\\,9 of Dorschner et al.\\ 1995, Fig.\\,1 of Voshchinnikov \\& Henning 2008). But for compact silicate spheres to have its 9.7$\\mum$ Si--O feature peaking at $\\lambda_{\\rm peak} >10.5\\mum$, their size needs to exceed $\\simali$2$\\mum$. However, for these large grains the 9.7$\\mum$ Si--O feature fades away (see Fig.\\,6 of Greenberg 1996, Fig.\\,1 of Voshchinnikov \\& Henning 2008). Therefore, it is difficult to account for the ``anomalous'' silicate emission features of AGNs just in terms of an increase in grain size. A non-spherical grain shape or shape distribution can also broaden the 9.7$\\mum$ Si--O feature and red-shift its peak wavelength (see Li 2008). But to account for the peak wavelengths of $\\lambda_{\\rm peak}$\\,$\\simali$10--11.5$\\mum$ observed in some AGNs, the dust has to be extremely elongated which is unrealistic (e.g. spheroidal dust needs to have an elongation $>$6). The peak wavelength of the 9.7$\\mum$ Si--O feature also depends on the silicate mineralogy. But we are not aware of any amorphous silicate species that could give rise to a Si--O feature peaking at $\\lambda >10\\mum$. Crystalline olivine has its Si--O feature peaks at $\\simali$11.2$\\mum$ (see Yamamoto et al.\\ 2008). But this feature is too sharp compared to the broad Si--O emission features of AGNs. Within the signal-to-noise ratio limits, we see no clear evidence for crystalline silicates in 3C\\,273 and NGC\\,3998 (see Fig.\\,\\ref{fig:sil_obs_spec}). So far, the detection of crystalline silicate dust in AGNs has only been reported in the BAL (broad absorption line) quasar PG\\,2112+059 (Markwick-Kemper et al.\\ 2007).\\footnote{% Spoon et al.\\ (2006) reported firstly the detection of narrow absorption features of crystalline silicates in ultraluminous IR galaxies (ULIRGs). } The 9.7$\\mum$ Si--O feature emission features of the protoplanetary disks around T Tauri and Herbig Ae/Be stars are often also much broader than that of the ISM (e.g. see Bouwman et al.\\ 2001, Forrest et al.\\ 2004). This is usually interpreted as an indicator of grain growth (e.g. see Natta et al.\\ 2007). The porous composite dust model was recently used by Voshchinnikov \\& Henning (2008) to demonstrate that a similar behavior of the feature shape occurs when the porosity of fluffy dust varies. Kr\\\"ugel \\& Siebenmorgen (1994) have also demonstrated that a porous structure results in the broadening, weakening and redshifting of the 9.7$\\mum$ Si--O feature. Fluffy dust aggregates are also naturally expected in protoplanetary disks as a result of grain coagulation. We plot in Figure \\ref{fig:agn_lambdap_porosity} as a function of porosity $P$ the peak wavelengths $\\lambda_{\\rm peak}$ of the 9.7$\\mum$ Si--O feature of the absorption efficiency profiles $\\Qabs(\\lambda)$ and the emission profiles obtained by folding $\\Qabs(\\lambda)$ with the Planck function $B_\\lambda(T)$ at various temperatures. It is seen that for 3C\\,273 and NGC\\,3998, to account for the observed $\\lambda_{\\rm peak}\\approx 10.6\\mum$, one requires either cool dust (with $T$\\,$\\simali$140--150$\\K$) with a low porosity ($P<0.3$), or warm dust (with $T$\\,$\\simali$170--220$\\K$) with a high porosity ($P>0.6$). The 18$\\mum$ O--Si--O emission feature further constrains that cool dust is preferred in 3C\\,273 while NGC\\,3998 appears to favour warm dust. Also seen in Figure \\ref{fig:agn_lambdap_porosity} is that the range of the peak wavelengths $10\\mum \\simlt \\lambda_{\\rm peak}\\simlt 11.5\\mum$ of the 9.7$\\mum$ Si--O emission features of the five PG quasars of Hao et al.\\ (2005) falls within the predictions of the porous composite dust model. This indicates that the long-wavelength shifted 9.7$\\mum$ Si--O features of these quasars can be accounted for by the combined effects of porosity and temperature (i.e. either highly porous warm dust or cold dust with a lower porosity). The 18$\\mum$ O--Si--O emission feature will allow us to break this degeneracy. Finally, for those AGNs whose 9.7$\\mum$ Si--O absorption or emission features do not exhibit any long-wavelength shift, one may resort to dust with a low porosity $P<0.3$ (and $T>220\\K$ if they are in emission). We should stress that although the combined effects of porosity and temperature suffice by themselves to explain the general trend of broadening the 9.7$\\mum$ Si--O emission features of AGNs and shifting their peak wavelengths to longer wavelengths, we by no means exclude the effects of other factors such as grain size,\\footnote{% The grain size effect is probably (at least in part) responsible for the nondetection of the 9.7$\\mum$ and 18$\\mum$ silicate emission features in many type 1 AGNs (e.g. see Hao et al.\\ 2007). The silicate size distribution in some AGNs might be dominated by large grains ($>$ a few $\\mum$; see Maiolino et al.\\ 2001a,b) or small silicate grains are depleted (e.g. see Laor \\& Draine 1993, Granato \\& Danese 1994). Alternative explanations include sophisticated torus geometries (e.g. tapered disk configurations [Efstathiou \\& Rowan-Robinson 1995], clumpy torus models [Nenkova et al.\\ 2002, 2008; but see Dullemond \\& van Bemmel 2005]) and an assumption of strong anisotropy of the source radiation (Manske et al.\\ 1998). } shape, and mineralogy. It is very likely that all these factors act together to produce the anomalous silicate emission features seen in AGNs. Finally, we emphasize that although a steeply rising (cold) Planck function could redshift the 9.7$\\mum$ Si--O emission feature, the observed shift of the peak wavelength of this emission feature in AGNs cannot be purely a temperature effect since the silicate absorption profiles of some AGNs also appear anomalous; e.g., the ``9.7$\\mum$'' silicate feature of Mkn 231, a peculiar type 1 Seyfert galaxy, is seen in absorption peaking at $\\simali$10.5$\\mum$ (Roche et al.\\ 1983); Jaffe et al.\\ (2004) found that the 9.7$\\mum$ silicate absorption spectrum of NGC\\,1068 shows a relatively flat profile from 8 to 9$\\mum$ and then a sharp drop between 9 and 10$\\mum$; in comparison, the Galactic silicate absorption profiles begin to drop already at $\\simali$8$\\mum$. To conclude, we have explored in a quantitative way on the effects of grain porosity on the silicate Si--O stretching feature using the multi-layered sphere model. It is found that the Si--O feature broadens and shifts to longer wavelengths with the increasing of dust porosity. We conclude that the combined effects of dust porosity and cool temperature of $T<$\\,200$\\K$ (which further redshifts the silicate feature) could explain the observed broadening and longer-wavelength shifting of the 9.7$\\mum$ Si--O feature of AGNs." }, "0808/0808.1575_arXiv.txt": { "abstract": "We report the discovery of a wide ($135\\pm25$~AU), unusually blue L5 companion 2MASS J17114559+4028578 to the nearby M4.5 dwarf G 203-50 as a result of a targeted search for common proper motion pairs in the Sloan Digital Sky Survey and the Two Micron All Sky Survey. Adaptive Optics imaging with Subaru indicates that neither component is a nearly equal mass binary with separation $> 0.18\\arcsec$, and places limits on the existence of additional faint companions. An examination of TiO and CaH features in the primary's spectrum is consistent with solar metallicity and provides no evidence that G 203-50 is metal poor. We estimate an age for the primary of 1-5 Gyr based on activity. Assuming coevality of the companion, its age, gravity and metallicity can be constrained from properties of the primary, making it a suitable benchmark object for the calibration of evolutionary models and for determining the atmospheric properties of peculiar blue L dwarfs. The low total mass ($M_{tot}=0.21\\pm0.03$~\\msun), intermediate mass ratio ($q=0.45\\pm0.14$), and wide separation of this system demonstrate that the star formation process is capable of forming wide, weakly bound binary systems with low mass and BD components. Based on the sensitivity of our search we find that no more than $2.2\\%$ of early-to-mid M dwarfs ($9.0 0.06$~\\msun. ", "introduction": "\\label{sect:intro} Although star birth is a complex process, the observation of binary systems--frequencies, mass ratios, and separations--can provide insight into the formation process as well as constraints for theoretical models. The formation of brown dwarfs (BDs) is particularly challenging since their masses are an order of magnitude smaller than the typical Jeans mass in molecular clouds. Whether BDs form similarly to their more massive stellar counterparts or require additional mechanisms is currently an open question. The answer may lie in the multiplicity properties of these substellar objects. Whereas free-floating BDs are observed in abundance, finding BDs as companions to stars has proved more difficult. A ``brown dwarf desert'' ($\\lesssim0.5\\%$ companion fraction) is observed at close separations ($<3$~AU) to main sequence stars, in comparison to a significant number of both planetary and stellar mass companions seen at similar separations \\citep{marcy00}. It has recently been determined that this desert does not extend out to larger separations for solar analogs (F,G,K stars), $\\sim7\\%$ of which are found to harbor substellar companions at separations greater than 30~AU \\citep{met05}. However, searches for substellar companions to M dwarfs at large separations ($\\gtrsim$ 40 AU) have yielded mostly null results \\citep[e.g.][]{allen08,mccarthy04,hinz02,daemgen07} or sparse results \\citep[e.g.][]{oppen01}. There are only 5 known BD companions to stellar M dwarf primaries at separations greater than 40AU: TWA 5b,c \\citep{lowrance99}, G 196-3B \\citep{rebolo98}, GJ 1001B \\citep{goldman99}, Gl 229B \\citep{nakajima95}, and LP 261-75B \\citep{burgasser05,reid06}. For thoroughness we note that the L2.5 companion GJ 618.1B \\citep{wilson01} may also fall into this category, however it is more likely stellar. In the VLM regime ($M_{1}<0.1$~\\msun) surveys have found that no more than $\\sim1\\%$ of stars have wide companions, including stellar ones \\citep{burgasser07a}. Additionally, VLM binaries are found to be on average 10-20 times more tightly bound than their stellar counterparts, hinting that disruptive dynamical interactions may play an important role in their formation \\citep{close03}. These observations have been cited as evidence in favor of the ejection hypothesis \\citep{rep01, bate05} where BDs and VLM stars are thought to be stellar embryos formed by the fragmentation of a more massive pre-stellar core, then prematurely ejected from their birth environments. However, BD companions to more massive stars do not tend to form harder binaries than stellar systems of similar total mass \\citep[e.g.][]{reid01,met05}. While this is potentially evidence that BDs can form similarly to stars via turbulent fragmentation within molecular clouds \\citep{padoan02}, it is also consistent with simulations of disk instabilities \\citep[e.g.][]{stamatellos07,boss00}, which are capable of producing substellar companions around more massive primaries. While a significant fraction of solar mass stars may retain wide BD companions, this does not seem to hold true for lower mass stars. As a result, very few wide BD companions to low mass stars are known. The discovery and characterization of these systems, especially in the intermediate range between the solar analog and VLM regimes, will help complete the emerging picture of BD multiplicity at wide separations. Additionally, wide BD companions to stars make suitable ``benchmark'' objects, as their properties can be inferred from those of the primary \\citep{pinfield06}. This is important for the calibration of BD evolutionary models, which requires independent age estimates. Here we present the discovery of a wide substellar companion to a nearby M4.5 star. The search, discovery and followup observations are outlined in \\S\\ref{sect:search}, while the physical properties of the system and its components are given in \\S \\ref{sect:properties}. In \\S\\ref{sect:discussion} we discuss the companion's unusual NIR colors, possible formation scenarios, and the sensitivity of our search. Given a space density for M dwarfs we make a crude estimate of how rare such systems may be. A brief summary and outlook are presented in \\S \\ref{sect:conclusions}. ", "conclusions": "\\label{sect:discussion} \\subsection{The blue NIR colors of 2M1711+4028}\\label{sect:blue} The NIR colors of mid-late L dwarfs vary significantly within a single spectral type. For L5 dwarfs there is a spread of $\\sim$0.7 magnitudes in $J-K_s$ \\citep{kirkpatrick08,cushing08}. Although surface gravity and metallicity play a role, comparison of atmospheric models to actual spectra \\citep[e.g.][]{knapp04, cushing08, burgasser08} suggests that large variations in the NIR colors of L dwarfs are primarily related to the properties of condensate clouds in their atmospheres, with unusually red SEDs arising from thick clouds and blue ones from thin or large-grained clouds. Common to the known peculiar blue L dwarfs is exaggerated H2O absorption and diminished CO, as seen in the spectrum of 2M1711+4028. As discussed in \\S \\ref{sect:secondary} the discrepancy between the late-type H20 indices and the earlier type FeH and K1 indices, along with its unusually blue NIR colors are indications that 2M1711+4028 falls into this category. For comparison we overplot 2M1711+4028's spectrum with the spectra of the very red L4.5 dwarf 2MASS J22244381-0158521 \\citep{kirkpatrick00}, and the relatively blue L5 optical standard 2MASS~J15074769-1627386 (see figure \\ref{fig:blue2}). All spectra agree reasonably well in the J-band but diverge significantly at H and K, which may indicate differing properties of condensate clouds in their atmospheres. As a member of a wide binary system the surface gravity and metallicity of 2M1711+4028 can be constrained from properties of the primary, yielding an excellent laboratory for studying BD atmospheres. The estimated age of 1-5~Gyr for the G 203-50 primary implies a relatively high surface gravity for the companion, possibly contributing the its blue NIR colors. However surface gravity alone does not seem to be sufficient in explaining the NIR colors of peculiar blue L dwarfs \\citep{burgasser08}. Additionally, the primary shows no signs of being metal poor, lending support to the hypothesis that unusually blue NIR colors can be primarily attributed to cloud properties. Higher resolution spectroscopy of the primary and a more precise determination of the metallicity is required in order to confirm this conclusion. Another potential cause of unusually blue NIR colors is unresolved binarity. This may explain the slight onset of CH4 absorption at 2.2 $\\mu$m in the spectrum of 2M1711+4028. At least one of the known peculiar blue L dwarfs, 2MASS J08053189+4812330 \\citep{burgasser07b} is thought to be an unresolved binary with L4.5 and T5 components. However, this system exhibits a pronounced dip in the 1.6 $\\mu$m CH4 feature due to the peaked shape of the T dwarf's SED, whereas the spectrum of 2M1711+4028 is relatively flat in that region. Furthermore, the SED and colors of our companion are very similar to that of another blue L4.5 dwarf, 2MASS J11263991-5003550 (2M1126-5003 hereafter), discussed at length by \\citet{burgasser08}. By constructing composite spectra using published L and T dwarf spectra from the SpeX prism library, \\citet{burgasser08} determined that no reasonable composite spectrum could be found that matched that of 2M1126-5003 in both the optical and NIR. Without an optical spectrum for 2M1711+4028 we are limited in the conclusions we can draw, but its similarities to 2M1126-5003 may suggest that 2M1711+4028 is a single BD. Our AO images support this conclusion, indicating that the BD is not a near-equal mass binary with separation $> 0.18\\arcsec$. \\subsection{Formation of G 203-50AB}\\label{sect:formation} With a total mass of $\\sim$0.21~\\msun~G~203-50AB is slightly more massive than the rare wide VLM binaries, but much less massive than the solar analogues around which BDs are routinely found at wide separations (see \\S \\ref{sect:intro}, and figure \\ref{fig:sep}). It is therefore of interest to consider how G~203-50AB may have formed. Could the secondary have formed through gravitational instability in a disk around the primary? Given the mass ratio of $q$=0.45, that would imply a $M_{disk} >$ 0.45$M_*$, whereas typical disks around low-mass stars contain a few percent of $M_*$ at $\\sim$1 Myr \\citep{scholz06}. Since it is unlikely that the entire disk would end up in the companion, the total disk mass, even in a conservative estimate, would have to be larger than the primary's own mass to start with. Thus, we conclude that formation of 2M1711+4028 in a protostellar disk around G~203-50 is implausible. On the other hand, gravitational fragmentation of prestellar cores appears to be capable of forming a wide variety of binary systems, depending on the size, mass and angular momentum of the core \\citep[e.g.][]{bate00}. However, simulations usually have some difficulty producing binary stars with low component masses and wide separations \\citep[e.g.][]{bate03,goodwin04}. Some theoretical models invoke ejection from the parent cloud to halt further accretion that would otherwise lead to higher masses. Given its projected separation of 135$\\pm$25~AU, the G~203-50AB binary has a binding energy of 12.6$\\pm3.8~\\times$~$10^{-41}$~erg, placing it below the empirical ``minimum'' noted by \\citet{close03} and \\citet{burgasser07a}. Thus, it is unlikely to have survived such an ejection. We suggest that G~203-50AB most likely formed via fragmentation of an isolated core and did not suffer strong dynamical interactions during the birth process or subsequently. \\subsection{Search Sensitivity}\\label{sect:sensitivity} In order to assess the sensitivity of our search we simulated proper motion distributions of M dwarfs distributed uniformly in a spherical volume out to 25 pc, with tangential space velocities drawn from the distribution of \\cite{schmidt07}. To be sensitive to a particular M dwarf primary, its displacement between the 2MASS and SDSS surveys had to be greater than the $3\\sigma$ dispersion of all other stars in the 4~${\\rm deg^2}$ section of the sky in which it was found. For each such section of the sky, the time baseline between the surveys was computed and used to determine the minimum proper motion required for a detection. We assumed that the population of M dwarfs within 25 pc was uniformly distributed and assigned equal weight to each 4 ${\\rm deg^2}$ area of the sky. Using the simulated proper motion distributions the fraction of M dwarfs whose proper motions we could have measured was determined for each section of the sky, giving an average fraction of 0.58. Adopting an M dwarf space density ($9.0 < M_V < 13.0$ or roughly M0.5-M5.5) of $283.37 \\times10^{-4}~{\\rm pc^{-3}}$ \\citep{reid02}, and given a search area of 7668 ${\\rm deg^2}$ of the sky, we should have been sensitive to approximately $201\\pm12$ early-mid M dwarfs within 25 pc. Although in some cases we were able to recover binary systems with separations $<$~$4\\arcsec$, we conservatively put a lower limit of $6\\arcsec$ on our sensitivity, ensuring that components are well separated. The upper limit for separation is set by our search radius, which extended to $120\\arcsec$. These limits correspond to projected separations of 30-600~AU at 5~pc and 150-3000~AU at 25~pc. Our sensitivity to companions around each star was dictated by the mean 2MASS J-band limiting magnitude (S/N=10) of $\\sim16.5$, corresponding to a minimum mass of $\\sim0.06$~\\msun~at 25~pc, assuming an age of 1-5 Gyr. Other factors preventing us from finding companions include poor astrometry due to saturation of the primary, or low S/N of the secondary. To estimate the number of binaries missed we used SIMBAD and DwarfArchives\\footnote{http://DwarfArchives.org} to compile a list of 31 M-dwarfs and 38 BDs with previously measured proper motions large enough to pass our cuts, and tested whether we could measure the same proper motions using SDSS and 2MASS astrometry. We found that $91\\%$ of the time for M dwarfs, and $79\\%$ of the time for BDs our measured proper motions agreed with the previously measured ones, using the same criteria as our matching algorithm described in \\S\\ref{sect:search}. Therefore, we should have been capable of identifying approximately $72\\%$ of binaries with sufficiently high proper motions. Correspondingly we adjust our sensitivity to $\\sim145\\pm9$ M dwarfs. Adopting Poisson uncertainties on a $1\\sigma$ confidence interval for our single detection we roughly estimate that $0.7^{+1.5}_{-0.6}\\%$ of early-mid M dwarfs have substellar companions with masses greater than $\\sim0.06$~\\msun, at separations above $\\sim120$~AU." }, "0808/0808.4159.txt": { "abstract": "\\noindent We investigate further a model of the accreting millisecond X-ray pulsars we proposed earlier. In this model, the X-ray--emitting regions of these pulsars are near their spin axes but move. This is to be expected if the magnetic poles of these stars are close to their spin axes, so that accreting gas is channeled there. As the accretion rate and the structure of the inner disk vary, gas is channeled along different field lines to different locations on the stellar surface, causing the X-ray--emitting areas to move. We show that this ``nearly aligned moving spot model'' can explain many properties of the accreting millisecond X-ray pulsars, including their generally low oscillation amplitudes and nearly sinusoidal waveforms; the variability of their pulse amplitudes, shapes, and phases; the correlations in this variability; and the similarity of the accretion- and nuclear-powered pulse shapes and phases in some. It may also explain why accretion-powered millisecond pulsars are difficult to detect, why some are intermittent, and why all detected so far are transients. This model can be tested by comparing with observations the waveform changes it predicts, including the changes with accretion rate. ", "introduction": "\\label{sec:intro} Highly periodic millisecond X-ray oscillations have been detected with high confidence in 22 accreting neutron stars in low-mass X-ray binary systems (\\mbox{LMXBs}), using the \\textit {Rossi X-ray Timing Explorer} (\\textit {RXTE}) satellite (see~\\citealt {lamb08a}). We refer to these stars as accreting millisecond X-ray pulsars (\\mbox{AMXPs}). Accretion-powered millisecond oscillations have so far been detected in 10 \\mbox{AMXPs}. They are always observable in seven \\mbox{AMXPs}, but are only intermittently detected in three others. Nuclear-powered millisecond oscillations have been detected with high confidence during thermonuclear X-ray bursts in 16 \\mbox{AMXPs}. Persistent accretion-powered millisecond oscillations have been detected in two \\mbox{AMXPs} that produce nuclear-powered millisecond oscillations; intermittent accretion-powered millisecond oscillations have been detected in two others. The \\mbox{AMXPs} have several important properties: \\textit {Low oscillation amplitudes}. The fractional amplitudes of the accretion-powered oscillations of most \\mbox{AMXPs} are often only $\\sim\\,$1\\%--2\\%.\\footnote{We characterize the strengths of oscillations by their rms amplitudes, because the rms amplitude can be defined for any waveform, is usually relatively stable, and is closely related to the power. We convert reported semi-amplitudes of purely sinusoidal oscillations or Fourier components to rms amplitudes by dividing by $\\sqrt2$.} Persistent accretion-powered oscillations with amplitudes $\\la\\,$1\\% are often detected with high confidence in IGR~J00291$+$5934 (\\citealt{gall05,patr08}) and XTE~J1751$-$305 (\\citealt{mark02, patr08}). Persistent accretion-powered oscillations with amplitudes as low as 2\\% are regularly seen in XTE~J1807$-$294 (\\citealt{zhan06, chou08, patr08}), XTE~J0929$-$314 (\\citealt{gall02}), and XTE~J1814$-$338 (\\citealt{chun08, patr08}). The amplitude of the accretion-powered oscillation seen in SWIFT~1756.9$-$2508 was $\\sim\\,$6\\% (\\citealt{krim07}). The intermittent accretion-powered oscillations detected in SAX~J1748.9$-$2021 (\\citealt{gavr07, alta08, patr08}), HETE~J1900.1$-$2455 (\\citealt{gall07}), and Aql~X-1 (\\citealt{case08}) have amplitudes $\\sim\\,$0.5\\%--3\\%. \\textit {Nearly sinusoidal waveforms}. The waveforms (light curves) of the accretion-powered oscillations of most \\mbox{AMXPs} are nearly sinusoidal (see \\citealt{wijn06} and the references in the preceding paragraph). The amplitude of the first harmonic (fundamental) component is usually $\\ga\\,$10 times the amplitude of the second harmonic (first overtone) component, although the ratio can be as small as $\\sim\\,$3.5, as is sometimes the case in XTE~J1807$-$294 (\\citealt{zhan06}), or even $\\ga\\,$1, as is sometimes the case in SAX~J1808.4$-$3658 (see, e.g., \\citealt{hart08}). \\textit {Highly variable oscillation amplitudes}. The fractional amplitudes of the accretion-powered oscillations of most \\mbox{AMXPs} vary in time by factors ranging from $\\sim\\,$2 to $\\sim\\,$10. Observed fractional amplitudes vary from 0.7\\% to 1.7\\% in SAX~J1748.9$-$2021, from 0.7\\% to 3.7\\% in XTE~J1751$-$305, from 3\\% to 7\\% in XTE~J0929$-$314, from 1\\% to 9\\% in IGR~J00291$+$5934, from 2\\% to 11\\% in XTE~J1814$-$338, from 1\\% to 14\\% in XTE~J1808$-$338, and from 2\\% to 19\\% in XTE~J1807$-$294 (see the references above). \\textit {Highly variable pulse phases}. The phases of accretion-powered pulses have been seen to vary rapidly by as much as $\\sim\\,$0.3 cycles in several \\mbox{AMXPs}, including SAX~J1808.4$-$3658 \\citep{morg03, hart08} and XTE~J1807$-$294 \\citep{mark04}. Wild changes in the apparent pulse frequency have been observed with \\textit {both} signs at the \\textit{same} accretion rate in XTE~J1807$-$294 (see \\citealt{mark04}). If interpreted as caused by changes in the stellar spin rate, these phase variations would be more than a factor of 10 larger than expected for the largest accretion torques and smallest inertial moments thought possible for these systems (see \\citealt{ghos79b, latt01}). \\textit {Undetected accretion-powered oscillations}. More than 80 accreting neutron stars in \\mbox{LMXBs} are known (\\citealt{chak05, liu07}), but accretion-powered millisecond X-ray oscillations have so far been detected in only 10 of them. Accretion-powered oscillations have not yet been detected even in 13 \\mbox{AMXPs} that produce periodic nuclear-powered millisecond oscillations, indicating that they have millisecond spin periods (\\citealt{lamb08a}); eight of these also produce kilohertz quasi-periodic oscillations (\\mbox{QPOs}) with frequency separations that indicate that they not only have millisecond spin periods but also have dynamically important magnetic fields (\\citealt{bout08a}). \\textit {Intermittent accretion-powered oscillations}. Accretion-powered millisecond X-ray pulsations have been detected only occasionally in SAX~J1748.9$-$2021 (\\citealt{gavr07, alta08, patr08}), HETE~J1900.1$-$2455 (\\citealt{gall07}), and Aql~X-1 (\\citealt{case08}). When oscillations are not detected, the upper limits are typically $\\la0.5$\\%. \\textit{Correlated pulse arrival times and amplitudes}. The phase residuals of the accretion-powered pulses of several \\mbox{AMXPs} appear to be anti-correlated with their fractional amplitudes, at least over some of the amplitude ranges observed. \\mbox{AMXPs} that show this type of behavior include XTE~J1807$-$294 and XTE~J1814$-$338 (\\citealt{patr08}). \\textit{Similar accretion- and nuclear-powered pulses}. The shapes and phases of the nuclear-powered X-ray pulses of the \\mbox{AMXPs} SAX~J1808.4$-$3658 (\\citealt{chak03}) and XTE~J1814$-$338 (\\citealt{stro03}) are very similar to the shapes and phases of their accretion-powered X-ray pulses. \\textit {Concentration in transient systems of \\mbox{AMXPs} with accretion-powered oscillations}. The \\mbox{AMXPs} in which accretion-powered oscillations have been detected tend to be found in binary systems that have outbursts lasting about a month (but see \\citealt{gall08}) separated by quiescent intervals lasting years (\\citealt{chak05, rigg08}). The accretion rates of these neutron stars are very low. In this paper we investigate further the ``nearly aligned moving spot'' model of \\mbox{AMXP} X-ray emission that we proposed previously \\citep{lamb06,lamb07,lamb08b}. This model has three main features: \\begin{enumerate} \\item The strongest poles of the magnetic fields of neutron stars with millisecond spin periods are located near---and sometimes very near---the stellar spin axis. This behavior is expected for several magnetic field evolution mechanisms. \\item The star's magnetic field channels accreting gas close to its spin axis, creating X-ray emitting areas there and depositing nuclear fuel there.\\footnote{\\cite{muno02} considered a single bright spot near the spin axis as well as a uniformly bright hemisphere and antipodal spots near the spin equator as possible reasons for the nearly sinusoidal waveforms of some X-ray burst oscillations, but did not consider accretion-powered oscillations or other consequences of emission from near the spin axis.} \\item The X-ray emitting areas on the stellar surface move, as changes in the accretion rate and the structure of the inner disk cause accreting gas to be channeled along different field lines to different locations on the stellar surface. (The magnetic field of the neutron star is fixed in the stellar crust on the timescales relevant to the phenomena considered here.) \\end{enumerate} These features provide the basic ingredients needed to understand the \\mbox{AMXP} properties discussed above. This is the subject of the sections that follow. As a guide to these sections, we summarize our results here. \\begin{enumerate} \\item Emission from near the spin axis naturally produces weak modulation, regardless of the viewing direction. The reason is that uniform emission from a spot centered on the spin axis is axisymmetric about the spin axis and therefore produces no modulation. Emission from a spot close to the spin axis has only a small asymmetry and therefore produces only weak modulation. \\item Emission from near the spin axis also naturally produces a nearly sinusoidal waveform, because the asymmetry of the emission is weak and broad. \\item If the emitting area is close to the spin axis, even a small movement in latitude can change the oscillation amplitude by a substantial factor. \\item If the emitting area is close to the spin axis, movement in the longitudinal direction by a small distance can change the phase of the oscillation by a large amount. Changes in the latitude and longitude of the emitting area are expected on timescales at least as short as the $\\sim\\,$0.1~ms dynamical time at the stellar surface and as long as the $\\sim\\,$10~d timescale of the variations observed in the mass accretion rate. \\item If the emitting area is very close to the spin axis and remains there, the oscillation amplitude may be so low that it is undetectable. The effects of rapid changes in the position of the emitting area---possibly in combination with other effects, such as reduction of the modulation fraction by scattering in circumstellar gas---may also play a role in reducing the detectability of accretion-powered oscillations in neutron stars with millisecond spin periods. These effects may explain the fact that accretion-powered X-ray oscillations have not yet been detected in many accreting neutron stars that are thought to have millisecond spin periods and dynamically important magnetic fields. \\item If the emitting area is very close to the spin axis, a small change in the accretion flow can suddenly channel gas farther from the spin axis, causing the emitting area to move away from the axis. This can make a previously undetectable oscillation become detectable. Temporary motion of the emitting area away from the spin axis may explain the intermittent accretion-powered oscillations of some \\mbox{AMXPs} (\\citealt{lamb09}). \\item If the pulse amplitude and phase variations observed in \\mbox{AMXPs} are caused by motion of the emitting area, they should be correlated. In particular, the pulse phase should be much more scattered when the pulse amplitude is very low. The reason is that changes in the longitudinal position of the emitting area by a given distance produce much larger phase changes when the emitting area is very close to the spin axis, which is also when the oscillation amplitude is very low. The observational consequences discussed so far depend only on features~(2) and~(3) of the model, i.e., that the accretion-powered X-ray emission of \\mbox{AMXPs} comes from areas near their spin axes and that these areas move significantly on timescales of hours to days. \\item The picture of \\mbox{AMXP} X-ray emission outlined here suggests that the shapes and phases of the nuclear- and accretion-powered pulses are similar to one another in some \\mbox{AMXPs} because the nuclear- and accretion powered X-ray emission comes from approximately the same area on the stellar surface. The reason for this is that in some cases, the mechanism that concentrates the magnetic flux of the accreting neutron star near its spin axis, as it is spun up, will naturally produce magnetic fields strong enough to confine accreting nuclear fuel near the magnetic poles at least partially, even though the dipole component of the magnetic field is weak. \\item The picture of neutron star magnetic field evolution and \\mbox{AMXP} X-ray emission outlined here also suggests a possible explanation for why the \\mbox{AMXPs} in which accretion-powered oscillations have been detected are in transient systems. If most neutron stars in \\mbox{LMXBs} were spun up by accretion from a low spin rate to a high spin rate, their magnetic poles were forced very close to their spin axes, making accretion-powered oscillations difficult or impossible to detect. However, those stars that are now in compact transient systems now experience infrequent episodes of mass accretion and the accretion rate is very low. By now they have been spun down from their maximum spin rates, a process that could force their magnetic poles away from their spin axes enough to produce detectable accretion-powered oscillations. \\end{enumerate} These last two observational consequences depend on feature~(1) of the model, i.e., on how the magnetic fields of neutron stars evolve as they are spun up and down by accretion and electromagnetic torques. In the remainder of this paper we discuss in detail the features of the model and its observational implications. In Section~\\ref {sec:model}, we outline our approach, discussing our modeling of X-ray emission from the stellar surface, our computational and the code verification methods, and the pulse profile representation we use. We present our results in Sections~\\ref {sec:amplitudes} and~\\ref {sec:variations}. These results are based on our computations of several hundred million waveforms for different emitting regions, beaming patterns, stellar models, and viewing directions. In Section~\\ref {sec:amplitudes}, we consider the shape and amplitude of X-ray pulses as a function of the size and inclination of the emitting areas, the compactness of the star, and the stellar spin rate. In Section~\\ref {sec:variations}, we consider the changes in the pulse amplitude and phase produced by various motions of the emitting regions on the stellar surface and explore the origins of correlated changes in the pulse amplitude and phase and the effects of rapid movement of the emitting areas. We also discuss why oscillations have not yet been detected in many accreting neutron stars in LMXBs and why accretion-powered oscillations are detected only intermittently in some \\mbox{AMXPs}. In Section~\\ref {sec:discussion}, we summarize the results of our model calculations. We also discuss how the magnetic poles of most \\mbox{AMXPs} can be forced close to their spin axes, how such mechanisms may explain why the \\mbox{AMXPs} that produce accretion-powered millisecond oscillations are transient pulsars, the consistency of the model with the observed properties of rotation-powered millisecond pulsars, and possible observational tests of the model discussed here. Further results of our investigation of the present model will be presented elsewhere (S. Boutloukos et al., in preparation). ", "conclusions": "\\label{sec:conclusions} In previous sections we have explored in some detail the nearly aligned moving spot model of \\mbox{AMXP} X-ray emission. In this model the X-ray emitting regions are close to the stellar spin axis and move around on the stellar surface with time. Here we list our principal conclusions. \\textit {Pulse amplitudes and shapes}. In Section~\\ref{sec:spot-inclination}, we investigated the amplitudes and shapes of the pulses produced by emitting spots on the stellar surface as a function of their inclination to the spin axis, for several X-ray beaming patterns and a range of stellar masses, compactnesses, and spin rates. We found that emitting areas on or near the stellar surface can produce fractional amplitudes as low as the 1\\%--2\\% values often observed only if they are located within a few degrees of the stellar spin axis. Regions near the spin axis also naturally produce nearly sinusoidal pulse profiles. We explored effect of spot size on pulse amplitude in Section~\\ref{sec:spot-size}. We found that although the pulse amplitudes produced by large emitting areas tend to be smaller, this effect is weak. Unless almost the entire surface of the star is uniformly emitting, even large spots produce pulse amplitudes greater than those observed in the \\mbox{AMXPs}, unless the spots are centered close to the spin axis. In Section~\\ref{sec:compactness}, we studied the effect of stellar compactness on pulse amplitudes. Although the pulse amplitudes produced by very compact neutron stars tend to be smaller than the amplitudes produced by less compact stars, we found that this effect is too weak to explain by itself pulse amplitudes as small as those observed in the \\mbox{AMXPs}. Stellar compactness clearly cannot explain why the fractional amplitudes of several \\mbox{AMXPs} are $\\sim\\,$1\\%--2\\% at some times but $\\sim\\,$15\\%--25\\% a few hours or days later, because the stellar compactness cannot change on such short timescales. These results show that emission from the stellar surface can explain the low amplitudes and nearly sinusoidal waveforms typically observed in \\mbox{AMXPs} only if the emitting areas are located close to the stellar spin axis. \\textit {Variability of pulse amplitudes, shapes, and arrival times}. In Section~\\ref{sec:amplitude-variations}, we investigated the amplitude changes that can be produced by motion of the emitting area on the stellar surface. We found that if the emitting area is close to the spin axis, even a small change in the latitude of the area can change the oscillation amplitude by a substantial factor. For example, changes in the inclination of the emitting area by $\\la\\,$10$\\arcdeg$ can explain the amplitude variations seen in the \\mbox{AMXPs} and the relatively large fractional amplitudes $\\sim\\,$15\\%--20\\% occasionally seen in some of them. In Section~\\ref{sec:phase-variations}, we studied the changes in the arrival times (phases) of the harmonic components of the pulse caused when the emitting area moves around on the stellar surface. We found that changes in the latitude of the emitting area can shift the phases of the first and second harmonics by at least 0.15~cycles and by different amounts. We showed that if the emitting area is close to the spin axis and moves in the azimuthal direction by even a small distance, the phases of the first and second harmonics can easily shift by as much as \\mbox{$\\sim\\,$0.1}--0.4 cycles. If the emitting area loops the spin axis, the phases of the Fourier components will shift by more than one cycle. Motion of the emitting area on the stellar surface generally produces both amplitude and phase variations. As discussed in Section~\\ref{sec:model}, the position of the emitting area is expected to reflect the accretion rate and structure of the inner disk, and is therefore expected to change on timescales at least as short as the $\\sim\\,$0.1~ms dynamical time at the stellar surface and as long as the $\\sim\\,$10~d timescale of the variations observed in the mass accretion rate. Changes in the position of the emitting area on timescales longer than the $\\sim\\,$$10^{2}$--$10^{3}$~s integration times required to construct a pulse profile will produce changes in the apparent pulse amplitude and phase. These results show that if the emitting area is close to the spin axis, modest changes in its location can explain the rapidly varying harmonic amplitudes and phases of the \\mbox{AMXPs}. \\textit {Correlated amplitude and phase variations}. We showed in Section~\\ref{sec:correlated-variations} that changes in the latitude and longitude of the emitting area tend to produce correlated changes in the amplitudes and phases of the harmonic components of the pulse. A strong expectation in the nearly aligned moving spot model is that the scatter in the pulse arrival times (i.e., the pulse time or phase residuals) should decrease steeply with increasing pulse amplitude. The residuals of several \\mbox{AMXPs}, including XTE~J1807$-$294, SAX~J1748.9-2021, and IGR~J00291$+$5934, behave in this way. The magnitudes of the phase residuals of these \\mbox{AMXPs} are consistent with the nearly aligned moving spot model if the emitting areas wander in the azimuthal direction by distances $\\sim\\,$$10^{-2}$--$10^{-3}$ times the stellar circumference. A second expectation in the nearly aligned moving spot model is that the arrival times of pulses will form a track in the phase-residual versus pulse-amplitude plane, especially if the change in the pulse amplitude is large. Such a track will be formed if the emitting areas where the accreting matter impacts the stellar surface move repeatedly along a particular path in stellar latitude and longitude as the accretion rate and the structure of the inner disk change. This is expected because models of the flow of accreting matter from the inner disk to the stellar surface predict that matter will be guided along particular but different magnetic flux tubes as the accretion rate and the structure of the inner disk vary (see Section~\\ref {sec:modeling-emission}). As discussed in Section~\\ref{sec:correlated-variations}, tracks of this type are observed in plots of pulse arrival time versus pulse amplitude for XTE~J1814$-$338 and XTE~J1807$-$294. \\textit {Accretion- and nuclear-powered oscillations}. The success of the emitting spot model discussed here supports magnetic field evolution models in which the magnetic flux of the accreting neutron star becomes concentrated near its spin axis as it is spun up. As discussed in Sections~\\ref{sec:accretion-nuclear} and~\\ref{sec:pole-movement}, these evolutionary models can produce magnetic fields strong enough to partially confine accreting nuclear fuel near the star's magnetic poles, even though the dipole component of the magnetic field is very weak. This picture of magnetic field evolution in turn suggests that the shapes and phases of the nuclear- and accretion-powered pulses are similar to one another in some \\mbox{AMXPs} because the nuclear- and accretion powered X-ray emission comes from approximately the same area on the stellar surface. \\textit {Effects of rapid spot motions}. In Section~\\ref{sec:spot-movements}, we discussed the effects of spot movements on timescales shorter than the time required to construct a pulse profile. Such effects are expected, because emitting spots are likely to move on the stellar surface on timescales at least as short as the $\\sim\\,$0.1~ms dynamical time there whereas constructing a pulse profile usually requires folding $\\sim\\,$$10^{2}$--$10^{3}$~s of X-ray flux data. Rapid spot motions will produce X-ray flux variations on these same timescales. Our computations show that motion of the emitting area on the stellar surface on timescales longer than the spin period usually changes the amplitudes and the phases of the harmonic components of the theoretical pulse profile. Variations of the X-ray flux on any timescales shorter than the time required to construct a pulse profile will appear in the analysis as noise in excess of the normal counting noise. This noise will reduce the measured amplitude of the oscillations at the spin frequency and its overtones for all spot locations, but its effect is likely to be stronger when the emitting area is near the spin axis because displacement of the emitting area by a given distance there produces a larger change in the pulse phase and, for many geometries, in the pulse amplitude. This effect will therefore tend to reduce the apparent pulse amplitude even further when the emitting area is close to the spin axis. \\textit {Undetectable and intermittent pulsations}. In Section~\\ref {sec:spot-inclination}, we showed that if the emitting areas of some \\mbox{AMXPs} are very close to the spin axis and remain there, the amplitudes of the oscillations they would produce can be $\\sim\\,$0.5\\% or less, making them undetectable with current instruments. Rapid X-ray flux variations will make accretion-powered oscillations at the spin frequency more difficult to detect. Other effects, such as scattering of X-ray photons in circumstellar gas, may also play a role in reducing the detectability of such oscillations. These results show that the nearly aligned moving spot model may, possibly in combination with other effects, explain the nondetection of accretion-powered oscillations at the millisecond spin frequencies of some accreting neutron stars in which nuclear-powered oscillations have been detected. In Section~\\ref {sec:amplitude-variations}, we showed that if the emitting area is within a few degrees of the spin axis and moves toward the rotation equator by $\\sim\\,$10$\\arcdeg$, oscillations that were undetectable can become detectable. This may explain why accretion-powered oscillations appear only intermittently in some \\mbox{AMXPs} \\citep{lamb08a}. The model suggests that oscillations may also disappear intermittently in some \\mbox{AMXPs}. \\textit {Evolution of \\mbox{AMXP} magnetic fields}. In Section~\\ref{sec:pole-movement}, we explained why the magnetic poles of most \\mbox{AMXPs} are expected to be very close to their spin axes. One consequence is that their magnetic fields will channel accreting gas to the stellar surface near the spin axis. Hence the star's accretion-powered X-ray emission will come from areas near the spin poles. A second consequence is that many \\mbox{AMXPs} may have surface magnetic fields as strong as $\\sim\\,$$10^{11}$--$10^{12}$~G, even though the dipole components of these fields are only $\\sim\\,$$10^{8}$--$10^{9}$~G. If so, many \\mbox{MRPs} may also have surface magnetic fields as strong as $\\sim\\,$$10^{11}$--$10^{12}$~G, even though the dipole components inferred from their spin-down rates are only $\\sim\\,$$10^{8}$--$10^{9}$~G. \\textit {Transient nature of \\mbox{AMXPs}}. In Section~\\ref{sec:transients}, we noted that the picture of \\mbox{AMXP} magnetic field evolution just described suggests why the \\mbox{AMXPs} in which accretion-powered oscillations have been detected are in transient systems. If the magnetic poles of most neutron stars in \\mbox{LMXBs} were forced very close to their spin axes during the initial, persistent phase of mass transfer, accretion-powered X-ray oscillations would be difficult or impossible to detect. Later, when mass transfer is transient, the stars will spin down, causing their magnetic poles to move away from their spin axes and making accretion-powered X-ray oscillations detectable. \\textit {Tests of the model}. The nearly aligned moving spot model leads to a number of expectations about \\mbox{AMXP} X-ray emission that can be tested (see Section~\\ref{sec:discussion} for details): \\begin{enumerate} \\item The amplitudes, harmonic content, and arrival times of pulses should be functions of the X-ray luminosity and spectrum of the pulsar. \\item The amplitudes, harmonic content, and arrival times of pulses should be correlated with one another. \\item The arrival times of pulses with low amplitudes should fluctuate much more than the arrival times of pulses with high amplitudes. \\item Pulse phase residuals are likely to form a track in the phase-residual versus pulse-amplitude plane. Such tracks are likely to be more evident if the range of the pulse amplitude variation is large. The position of pulses along such a track should be correlated with the X-ray luminosity and spectrum of the pulsar. \\item If the accretion- and nuclear-powered pulses of an \\mbox{AMXP} appear nearly identical, long-term (days-to-weeks) variations in the phase residuals of the two types of pulses should track one another. \\item The strength of the excess background noise produced by pulse shape fluctuations should be correlated with the amplitudes, harmonic content, and arrival times of pulses and the X-ray luminosity and spectrum of the pulsar. \\item The excess noise produced by motion of the emitting area should be stronger when the pulse amplitude is smaller and weaker when the pulse amplitude is larger. \\item If \\mbox{AMXPs} do have surface magnetic fields as strong as $\\sim\\,$$10^{11}$--$10^{12}$~G, their keV X-ray spectra may show strong-magnetic-field features. Such features are more likely in \\mbox{AMXPs} that show evidence of nuclear fuel confinement, such as SAX~J1808.4$-$3658 and XTE~J1814$-$338. \\item If \\mbox{AMXPs} have surface magnetic fields $\\sim\\,$$10^{11}$--$10^{12}$~G, their offspring should have millisecond spin periods and total magnetic fields of similar strength, but dipole fields $\\sim\\,$$10^{8}$--$10^{9}$~G. This should affect particle acceleration and $\\gamma$-ray emission by these neutron stars. \\item If the explanation of the transient nature of the \\mbox{AMXPs} suggested here is correct, they should be spinning down on long timescales. \\end{enumerate} \\textit{Note added in manuscript:} This work was presented at the 2008 April Amsterdam Workshop on Accreting Millisecond Pulsars, where we emphasized a test of our nearly aligned moving spot model that could also have wider implications. We argued that the close similarity of the accretion-powered and burst oscillation waveforms in XTE~J1814$-$338 strongly suggests that the emitting regions that produce them are similar and collocated, and that if, as we had proposed previously, the phase wandering of the accretion-powered oscillations is caused by movement of the emitting region, then the phase of the burst oscillations should wander in the same way. After the Workshop, \\citet{watt08} investigated this possibility and found just such a correlation in the \\textit {RXTE} data on XTE~J1814$-$338. In particular, these authors found that during the main part of the two-month outburst of XTE~J1814$-$338, its burst oscillation not only has a waveform similar to that of the accretion-powered oscillation \\citep{stro03, watt05, watt06}, but is also phase-locked with it, and that the peak of the burst oscillation coincides with the peak of the soft component of the accretion-powered oscillation. This indicates that, as we had suggested, the accretion- and nuclear-powered emitting regions in this pulsar very nearly coincide, and that the simultaneous wandering of the arrival times of both oscillations by $\\sim\\,$1~ms ($\\sim\\,$0.3 in phase) during the outburst is due to wandering of the matter (and hence the fuel) deposition pattern on the stellar surface." }, "0808/0808.0471.txt": { "abstract": "We have performed a census of circumstellar disks around brown dwarfs in the $\\sigma$~Ori cluster using all available images from the Infrared Array Camera onboard the {\\it Spitzer Space Telescope}. To search for new low-mass cluster members with disks, we have measured photometry for all sources in the {\\it Spitzer} images and have identified the ones that have red colors that are indicative of disks. We present five promising candidates, which may consist of two brown dwarfs, two stars with edge-on disks, and a low-mass protostar if they are bona fide members. Spectroscopy is needed to verify the nature of these sources. We have also used the {\\it Spitzer} data to determine which of the previously known probable members of $\\sigma$~Ori are likely to have disks. By doing so, we measure disk fractions of $\\sim40$\\% and $\\sim60$\\% for low-mass stars and brown dwarfs, respectively. These results are similar to previous estimates of disk fractions in IC~348 and Chamaeleon~I, which have roughly the same median ages as $\\sigma$~Ori ($\\tau\\sim3$~Myr). Finally, we note that our photometric measurements and the sources that we identify as having disks differ significantly from those of other recent studies that analyzed the same {\\it Spitzer} images. For instance, previous work has suggested that the T dwarf S~Ori~70 is redder than typical field dwarfs, which has been cited as possible evidence of youth and cluster membership. However, we find that this object is only slightly redder than the reddest field dwarfs in $[3.6]-[4.5]$ ($1.56\\pm0.07$ vs.\\ 0.93--1.46). We measure a larger excess in $[3.6]-[5.8]$ ($1.75\\pm0.21$ vs.\\ 0.87--1.19), but the flux at 5.8~\\micron\\ may be overestimated because of the low signal-to-noise ratio of the detection. Thus, the {\\it Spitzer} data do not offer strong evidence of youth and membership for this object, which is the faintest and coolest candidate member of $\\sigma$~Ori that has been identified to date. ", "introduction": "\\label{sec:intro} The lifetime of an accretion disk around a young star represents a fundamental constraint on the amount of time available for the formation of giant planets. The typical lifetimes of disks are estimated by comparing the prevalence of disks among clusters that span a range of ages ($\\tau\\sim1$--10~Myr). By measuring disk fractions as a function of stellar mass and star-forming environment as well as age, the influence of these two factors on disk lifetimes can be characterized. Disk fractions are usually measured through infrared (IR) photometry of young clusters ($\\lambda>3$~\\micron) and the identification of the stars that exhibit excess emission from cool dust. Observations of this kind were first performed with the {\\it Infrared Astronomical Satellite} and ground-based near-IR cameras \\citep{kh95,hai01} and have made rapid progress in recent years with the {\\it Spitzer Space Telescope} \\citep{wer04}. Because {\\it Spitzer} is exceptionally well-suited for detecting disks around members of young clusters \\citep{all04,gut04,meg04,muz04}, it has been used extensively for measuring the disk fractions of stars \\citep{meg05,lada06,car06,sic06,her07,her07b,her08,mue07,dahm07,bar07,dam07,luh08cha}. {\\it Spitzer}'s unique sensitivity to faint IR sources has enabled measurements of disk fractions well into the substellar regime in nearby star-forming regions \\citep[$d=150$--300~pc,][]{luh05frac,luh06tau2,luh08cha,gui07}. The cluster of young stars associated with the O star $\\sigma$~Ori~A has been thoroughly searched for low-mass stars and brown dwarfs \\citep[e.g.,][]{bej01,bar02,mar01,zap02a,zap02b}. As a result, it is an attractive target for measuring the disk fraction among low-mass objects. Several recent studies have sought to do this with {\\it Spitzer}. \\citet{her07} performed {\\it Spitzer} imaging of most of the $\\sigma$~Ori cluster and measured the disk fraction as a function of mass down to $\\sim0.1$~$M_\\odot$ for a sample of $\\sim300$ probable members (see also \\citet{cab06}). \\citet{cab07} identified a sample of brown dwarf candidates from optical and near-IR data and used the {\\it Spitzer} images from \\citet{her07} to estimate a disk fraction for those sources, arriving at a value of $\\sim50$\\%. Through further analysis of those {\\it Spitzer} data, \\citet{zap07} found that 6 of 12 brown dwarf candidates exhibited excess emission at 8~\\micron. \\citet{sch08} obtained deeper {\\it Spitzer} images of 18 of the faintest candidate members and reported a disk fraction of 29\\%. In addition to measuring disk fractions, the {\\it Spitzer} data have been used to assess the youth and membership of the coolest candidate member of the cluster, the T dwarf S~Ori~70 \\citep{zap02a}. \\citet{zap08} found that it is redder than typical field dwarfs in $[3.6]-[4.5]$ in the images from \\citet{her07}. Similarly, \\citet{sch08} suggested that S~Ori~70 could be anomalously red in $[3.6]-[4.5]$ and $[3.6]-[5.8]$ based on their deeper images, which they attributed to a disk or low surface gravity. In either case, the apparent color excesses would comprise evidence of youth for this object, whose membership in the $\\sigma$~Ori cluster has been questioned \\citep{mar03,bur04}. The substellar population of the $\\sigma$~Ori cluster extends down to and below the detection limits of the images that have been obtained by {\\it Spitzer}. Thus, determining whether these objects have disks is a challenging task. For instance, \\citet{zap07} and \\citet{sch08} disagree on whether the {\\it Spitzer} data show evidence of disks for half of the sources considered by both studies. In this paper, we seek to accurately characterize the disk population among low-mass stars and brown dwarfs in $\\sigma$~Ori through an analysis of all {\\it Spitzer} images of this cluster that includes careful treatment of errors and biases. We begin by summarizing the {\\it Spitzer} observations and our data reduction methods (\\S~\\ref{sec:obs}). We then use these data to search for new low-mass members of $\\sigma$~Ori that have disks (\\S~\\ref{sec:select}) and to estimate the disk fraction for the substellar members of the cluster (\\S~\\ref{sec:pop}). ", "conclusions": "\\label{sec:conc} We have investigated the disk population among stars and brown dwarfs in the $\\sigma$~Ori cluster using mid-IR images obtained with IRAC onboard the {\\it Spitzer Space Telescope}. For this study, we have employed all available IRAC data for $\\sigma$~Ori, which consist of shallow images (80.4~s) from \\citet{her07} and deep images ($\\sim1100$~s) from \\citet{sch08}. We measured photometry for all sources detected in these images and searched the resulting data for new members of the cluster based on the red colors that are expected from stars with disks. The five most promising candidates have colors and magnitudes that are suggestive of edge-on disks, brown dwarfs, and a low-mass protostar. We then examined the IRAC colors for $\\sim300$ probable cluster members found in previous studies and identified the ones that are likely to have disks. In doing so, we have attempted to fully account for the errors and biases in the photometry of the brown dwarf candidates \\citep[e.g., flux overestimation at low SNR;][]{bei03}, some of which are near the detection limits of the IRAC data. S~Ori~60 is the faintest candidate member of $\\sigma$~Ori that exhibits significant IR excess emission. This object is comparable in luminosity to the faintest brown dwarf that shows evidence of a disk in other star-forming regions \\citep{luh08cha}. Using our classifications of the IRAC data, we computed the disk fraction as a function of $M_J$, which acts as a proxy for stellar mass. The disk fractions for low-mass stars (0.08--0.5~$M_\\odot$) and brown dwarfs (0.01--0.08~$M_\\odot$) are $\\sim40$\\% and $\\sim60$\\%, respectively, which are similar to the disk fractions derived from IRAC surveys of two other clusters near the same age, IC~348 and Chamaeleon~I \\citep{mue07,luh08cha}. Although our disk fraction for brown dwarfs in $\\sigma$~Ori is similar to other estimates based on the IRAC data considered here \\citep{cab07,zap07,sch08}, our photometric measurements and our classifications of the IRAC colors (disk vs.\\ no disk) differ significantly from those in previous studies. For instance, \\citet{zap08} and \\citet{sch08} suggested that the T dwarf S~Ori~70 may exhibit color excesses in $[3.6]-[4.5]$ and $[3.6]-[5.8]$ relative to field dwarfs at the same spectral type. They cited these non-standard colors as possible evidence of youth, which would indicate that S~Ori~70 is a cluster member rather than a field dwarf. However, we have found that this object is not significantly redder in $[3.6]-[4.5]$ than field dwarfs and that its SNR may be too low at 5.8~\\micron\\ for a useful measurement of $[3.6]-[5.8]$. Therefore, we conclude that the IRAC data do not provide firm constraints on the membership of S~Ori~70." }, "0808/0808.3436_arXiv.txt": { "abstract": "We observed the brightest part of HESS J1825$-$137 with the Suzaku XIS, and found diffuse X-rays extending at least up to $\\timeform{15'}$ ($\\sim 17$~pc) from the pulsar PSR~J1826$-$1334. The spectra have no emission line, and are fitted with an absorbed power-law model. The X-rays, therefore, are likely due to synchrotron emission from a pulsar wind nebula. The photon index near at the pulsar ($r\\leq1.5'$) is $\\Gamma=1.7$ while those in $r=1.5-16'$ are nearly constant at $\\Gamma=2.0$. The spectral energy distribution of the Suzaku and H.E.S.S. results are naturally explained by a combined process; synchrotron X-rays and $\\gamma$-rays by the inverse Compton of the cosmic microwave photons by high-energy electrons in a magnetic field of $\\sim 7~\\mu$G. If the electrons are accelerated at the pulsar, the electrons must be transported over 17~pc in the synchrotron life time of 1900~yr, with a velocity of $\\geq 8.8 \\times 10^3$~km~s$^{-1}$. ", "introduction": "PSR~J1826$-$1334 was discovered by the Jodrell Bank radio survey. The spin period ($P$) is 101~ms and the period derivative ($\\dot{P}$) is $7.5\\times10^{-14}~\\mathrm{s~s^{-1}}$, implying a spin-down luminosity ($\\dot{E}$) of $2.8 \\times 10^{36} \\mathrm{~erg~s^{-1}}$ and a characteristic age ($\\tau$) of 21~kyr~\\citep{Clifton1992}. The distance to PSR~J1826$-$1334 has been measured as $3.9 \\pm 0.1$~kpc by the dispersion measure of its radio pulses~\\citep{Taylor1993,Cordes2002}. From the spin-down luminosity and characteristic age, PSR~J1826$-$1334 may be a member of the Vela-like pulsars, in the evolutionary stage of the high-energy emission processes dominance. There are pulsed non-thermal and blackbody emission from a pulsar, an extended nebulosity powered by pulsar wind (Pulsar Wind Nebula; PWN), and thermal and/or non-thermal emission from a supernova remnant (SNR)~\\citep{Bec97,Gaensler2003PSRB1823}. Extended X-rays around PSR~J1826$-$1334 had been found with the ROSAT~\\citep{Finley1996} and ASCA~\\citep{Sakurai2001} observations, possibly due to PWN, but no extended radio emission was found ~\\citep{Bra89,Gae00}. XMM-Newton resolved the extended X-rays into two parts; \"core\", a bright and elongated emission in the east-west direction with an extent of \\timeform{30\"}, and \"diffuse\", a larger scale with lower surface brightness emission extending by $\\sim \\timeform{5'}$ mostly to the south of the pulsar \\citep{Gaensler2003PSRB1823}. Spectra of both components can be fitted with an absorbed power-law model, and hence are suggested to be a synchrotron origin. The photon index of the core and diffuse components of $\\Gamma \\sim 1.6$ and harder $\\Gamma$ of $\\sim 2$, respectively. The absorbing column densities $N_{\\mathrm{H}}$ are both about $1 \\times 10^{22}~\\mathrm{cm^{-2}}$. \\citet{Gaensler2003PSRB1823} interpreted both the components as parts of a single PWN and designated as G18.0$-$0.7. They argued that the non-detection of the PWN in the radio band is due to the limited sensitivity for a nebula of this large size. The one-sided morphology of the diffuse component suggests that the PWN is extended exclusively to the south of the pulsar. The Chandra observation~\\citep{Pavlov2008} confirmed the existence of the two components, and measured the X-ray spectrum of the pulsar itself for the first time. The photon index of the pulsar is $\\Gamma \\sim 2.4$, and the luminosity is $8 \\times 10^{31}~\\mathrm{erg~s^{-1}}$ in the 0.5--8 keV band. HESS~J1825$-$137 is one of the handful Very High Energy (VHE) $\\gamma$-ray sources on the Galactic plane discovered during the H.E.S.S. Galactic plane survey~\\citep{Aharonian2005a, Aharonian2005b, Aharonian2006Survey}. It is extending to the south of the radio pulsar PSR~J1826$-$1334. Subsequently, deep follow-up observation was made with H.E.S.S. ~\\citep{Aharonian2006HESSJ1825} The $\\gamma$-ray emission is largely extended to $\\sim \\timeform{1D}$, i.e. $\\sim 70$~pc in size at the distance of 4~kpc. It has a larger scale one-sided morphology than X-rays toward the south of the pulsar. However, the detailed morphology is different from the X-ray; the surface brightness decreases and $\\gamma$-ray spectrum shows softening with the distance from the pulsar. These suggest that the HESS~J1825$-$137 is a PWN powered by PSR~J1826$-$1334. The most probable scenario for the VHE $\\gamma$-ray emission is inverse Compton (IC) scattering of the cosmic microwave background (CMB) photons with ultra-relativistic electrons accelerated by the pulsar, which is also responsible for the X-ray emission via the synchrotron process. However, the VHE $\\gamma$-rays are more extended than the X-rays. Furthermore, the X-rays are weak at the peak position of the $\\gamma$-rays. If the $\\gamma$-rays have the electron origin (IC), the typical energy of electrons is $\\sim$20~TeV. On the other hand, the typical energy of electrons emitting synchrotron X-rays is $\\sim$100~TeV assuming a magnetic field of $B=10~\\mu$G. These apparent differences between X-rays and $\\gamma$-rays may play a key role for the PWNe physics, such as the energy transport mechanism of and the history of energy injections to electrons. The extent of the diffuse X-ray emission, however, may be underestimated, due mainly to the limited sensitivity XMM-Newton and Chandra for diffuse X-rays. We hence observed the PWN with the X-ray Imaging Spectrometers~\\citep{Koyama2007XIS} on board the Suzaku satellite~\\citep{Mitsuda2007Suzaku}. In this paper, we will introduce the observations in section 2, our analysis and results are explained in section 3, and we will discuss our results in section 4. Uncertainties are quoted at the 90\\% confidence range unless otherwise stated. ", "conclusions": "\\subsection{Radial Profile} In figure~\\ref{fig:radialprofile}, we see largely extended emission up to \\timeform{15'} for the first time. The radial profile becomes flat beyond \\timeform{10'}. The flux in this flat region is 50\\% larger than the mean flux of the background region. One may argue that the difference could be due to the fluctuations of the NXB, CXB and GRXE. We therefore discuss whether any systematic errors of the background subtraction (NXB, CXB and GRXE) may mimic such excess. The reproductively of the NXB is extensively examined by \\citet{Tawa2008}. Using the same method, the NXB reproductively in the 1.0--9.0 keV band at the 99\\% confidence level is less than $\\mathrm{2.2 \\times10^{-5}~counts~s^{-1}~arcmin^{-2}}$. This uncertainty is very small compared with the excess level of the radial profile of the source region (figure~\\ref{fig:radialprofile}). The fluctuation of the CXB in the XIS FOV is estimated to be $\\sim$ 36\\% at the 99\\% confidence level with the same way as \\citet{Tawa2008}. Since the CXB flux is only 10\\% of the background spectrum, the overall uncertainty is less than 4\\%. The largest uncertainty would come from the subtraction of the GRXE. \\citet{Yam93} and \\citet{Kan97} reported that the GRXE has a large scale structure depending on the galactic longitude ($l$) and latitude ($b$). The $l$-dependency is rather null in this region hence can be neglected in our case. The $b$-dependency is expressed by the scale height of the medium and hard components given by 2 and 0.7 degrees, respectively \\citep{Yam93, Kan97}. Since the flux of the medium and hard components in our region is nearly the same in the 1.0-9.0 keV band, we can estimate that GRXE flux at the source position is about 20\\% larger than that in the background position. The most unknown factor is a possible fluctuation of the GRXE from position to position, which may cause under/over subtraction of the GRXE. If there is under/over subtraction of the GRXE, the source spectra should have emission/absorption line structures at the Si, S and Fe K-shell transition energies, because the GRXE has strong lines of these elements (see figure \\ref{fig:spectrumBG}). No such line/absorption structures are found in the spectra of the regions B, C, and D given in figure 5, which supports that the GRXE is properly subtracted. We further checked possible under/over subtraction of the GRXE by fitting the source spectrum (combined B+C+D) with a model of absorbed power-law + $\\Delta\\times$ GRXE, where the spectral parameters of the GRXE, other than normalization ($\\Delta$) were fixed to the best-fit values given in table~\\ref{tab:BGfitting}. As a result, we can set the 90\\% upper-limit of the possible contamination of the GRXE to be 1.5\\% of the excess flux. Accordingly, the largely extended power-law component from the source region is real, not an artifact of improper GRXE subtraction. \\subsection{Comparison with earlier X-ray observation} In the XMM-Newton observation, \\citet{Gaensler2003PSRB1823} reported that the photon indices of the core and diffuse components are $\\sim$ 1.6 and $\\sim$ 2, respectively. Since the region A includes both the components, our result of $\\Gamma \\sim 1.7$ is reasonable. From the region B, we determined more accurate photon index and interstellar absorption than the diffuse component of XMM-Newton. Moreover, we detected the X-ray emission farther out of the previously reported region. The peak position of the extended emission coincides with the pulsar PSR~J1826$-$1334, and the surface brightness decreases with the distance from the pulsar. The major fraction of the extended emission cannot be a thin thermal plasma of the SNR, but is non-thermal origin. The photon indices are smaller than the canonical value of synchrotron emissions found in non-thermal SNRs (2.5--3.0; Bamba~et al.~2005). Furthermore, the morphology of the extended emission is different from limb-brightened structures such as SN1006~\\citep{Koyama1995}. Thus the extended emission cannot be explained as non-thermal emission from shell-like SNR. The photon indices are close to typical values of PWNe~\\citep{Gaensler2006}. Therefore, we conclude that the extended emission is synchrotron radiation from a PWN, as suggested first by \\citet{Gaensler2003PSRB1823}. Our results indicate that the PWN extends over $\\timeform{15'}$, which corresponds to 17$d_\\mathrm{4kpc}$~pc, where $d_\\mathrm{4kpc}$ is the distance of PSR~J1826$-$1334 normalized by 4~kpc. The interstellar absorption of $N_{\\mathrm{H}} \\sim 10^{22} \\mathrm{cm}^{-2}$ is consistent with the distance of $\\sim 4$~kpc obtained by the radio measurements~\\citep{Taylor1993,Cordes2002}. The extent of $\\sim 17d_{\\rm 4kpc}$~kpc is larger than typical X-ray PWNe; for example, one of the largest PWNe is 3C58, where the X-ray nebula extends up to $\\sim 6$~pc from the pulsar~\\citep{Slane2004}. The total X-ray luminosity ($L_{\\rm X}$) from the regions A--D are $8.6 d_{\\rm 4kpc}^2 \\times 10^{33}$~erg~s$^{-1}$ in the 2--10 keV band. The ratio of $L_{\\rm X}$ to the spin-down luminosity is not unusual compared with other PWNe~\\citep{Che04,Li07}. \\subsection{Comparison with $\\gamma$-ray observation} Since the region C overlaps the bright region of the VHE $\\gamma$-rays, we can make a reliable spectral energy distribution (SED) between the X-rays and the VHE $\\gamma$-rays for the first time. In figure~\\ref{fig:SED}, we show the X-ray SED of the region C and the $\\gamma$-ray SED of the brightest region (Radius \\timeform{0.15D} in table~2 of \\cite{Aharonian2006HESSJ1825}). If the origin of the VHE $\\gamma$-rays is the inverse Compton scattering of the CMB photons by high-energy electrons, these electrons should emit synchrotron radiation. We also plot the estimated synchrotron intensity in the magnetic fields of $B= 10,~5,~1~\\mu$G. In the case of $B \\sim 7~\\mu$G, the estimated synchrotron intensity smoothly connects to the X-ray intensity. Thus both the X-rays and VHE $\\gamma$-rays can be explained by high-energy electrons of a single population in a magnetic field of $B \\sim 7~\\mu$G, approximately the same as the core region of $B \\sim 10~\\mu$G ~\\citep{Gaensler2003PSRB1823}. \\begin{figure} \\begin{center} \\FigureFile(80mm,60mm){figure6.eps} \\end{center} \\caption{ Spectral energy distribution of the extended component from the X-ray to TeV $\\gamma$-ray bands. The intensity of the region C in table~\\ref{tab:fittings} and that of the Radius \\timeform{0.15D} region in table~2 of \\citet{Aharonian2006HESSJ1825} are used for X-ray and VHE $\\gamma$-ray. The synchrotron radiation from the electrons responsible for VHE $\\gamma$-ray (0.25$-$10~TeV) is plotted toward the left for three different values of the magnetic field. \\label{fig:SED}} \\end{figure} According to \\citet{deJager2008}, required electron energy $E_\\mathrm{e}$ in a magnetic field $B$ to radiate synchrotron photons of mean energy $E_\\mathrm{syn}$ is given by $E_\\mathrm{e}$ $=$ $120~\\mathrm{TeV}~(B/7\\mu \\mathrm{G})^{-1/2} (E_\\mathrm{syn} / \\mathrm{2keV})^{1/2}$. On the other hand, the mean electron energy $E_\\mathrm{e}$ required to inverse Compton scatter the CMB photons to energies $E_\\mathrm{IC}$ is typically $E_\\mathrm{e}=18~\\mathrm{TeV}~(E_\\mathrm{IC} /\\mathrm{1TeV})^{1/2}$ \\citep{deJager2008}. Thus electrons responsible for X-rays have larger energies than those for VHE $\\gamma$-rays. The peak position of the VHE $\\gamma$-rays does not coincide with the pulsar~\\citep{Aharonian2006HESSJ1825}, while the peak position in X-rays are consistent with the pulsar. There is no indication of the $\\gamma$-ray peak in the XIS images (figures~\\ref{fig:image} and \\ref{fig:radialprofile}). The offset is a mystery, and the reason is not yet clarified. However, as \\citet{Pavlov2008} mentioned, the local excess of the seed photons for the inverse Compton scattering created by the near hidden star might cause the offset. In this case, a magnetic field stronger than $\\sim 10~\\mu$G might be required. \\citet{Har99} reports an EGRET source 3~EG J1826--1302 in the north of PSR~J1826--1334, and the relation between them has been discussed by some authors (e.g. Zhang, Chen \\& Fang~2008). The EGRET source is , however, out of the FOV of the XIS in our observation. Thus direct comparison between the intensities of the EGRET source and the X-ray we found is impossible, and the relation between them is not clarified. \\subsection{Widely extended PWN} Except for the neighborhood of the pulsar, the photon index is constant at $\\Gamma=2.0$ from the pulsar to $\\timeform{15'}$. The smooth decay of the intensity with the distance from the pulsar shown in figure~\\ref{fig:radialprofile} suggests that the accelerator is the pulsar itself. This means the high-energy electrons should be transported over $17d_\\mathrm{4kpc}$~pc within its synchrotron life time. According to \\citet{deJager2008}, the synchrotron life time of an electron emitting photons of mean energy $E_\\mathrm{syn}$ in an isotropic magnetic field of the strength $B$ is $\\tau_\\mathrm{syn}$ $=$ $1.9~{\\rm kyr}~(B/7\\mu \\mathrm{G})^{-3/2} (E_\\mathrm{syn} / \\mathrm{2keV})^{-1/2}$. Thus, the velocity for transporting the high-energy electrons should be more than $17 d_\\mathrm{4kpc}$~pc$/ \\tau_\\mathrm{syn}$ $\\sim$ $8.8 \\times 10^3~{\\rm km~s}^{-1} d_\\mathrm{4kpc} (B/7\\mu \\mathrm{G})^{3/2} (E_\\mathrm{syn} / \\mathrm{2keV})^{1/2}$. A simple transport mechanism is diffusion by a magnetic field or convection. In the case of diffusion, our results indicate that the diffusion coefficient is $D=R^2/(2\\tau) = 2.3 \\times 10^{28} \\mathrm{~cm^2~s^{-1}} (R/\\mathrm{17pc})^2 (\\tau/1.9\\mathrm{kyr})^{-1}$, where $R$ is the size of the extended emission, and $\\tau$ is the transport time for which we assume the synchrotron life time. We can express the mean free path of the electron in a magnetic field $B$ as $f r_\\mathrm{L}$, where $r_\\mathrm{L} = E_e/(eB)$ is the gyro radius of an electron with energy $E_e$ and charge $e$, and the parameter $f$ characterizes the efficiency of diffusion. Then the diffusion coefficient can also be expressed as $D$ $\\sim$ $f r_\\mathrm{L} c /3$ $=$ $2.3\\times 10^{28} \\mathrm{~cm^2~s^{-1}}~(B/7\\mu \\mathrm{G})^{-1}~(E_e/120\\mathrm{TeV}) (f/40)$. Thus, our results suggest $f \\sim 40$. According to \\citet{deJager2008}, $f$ should be $\\le 1$ in perpendicular diffusion. Therefore, our results suggest that the transport mechanism is not the perpendicular diffusion, but it might be diffusion parallel to the magnetic field or convection. Collisions between a PWN and a reverse shock from the surrounding SNR were proposed as a scenario to explain the morphology of one-sided PWNe \\citep{Pavlov2008}, such as Vela pulsar~\\citep{Blondin2001, Gaensler2003PSRB1823}. Whether or not this scenario can explain the large extent of $\\sim$17$d_{\\rm 4kc}$~pc and the smooth radial profile shown in figure~\\ref{fig:radialprofile} is an open problem." }, "0808/0808.3957_arXiv.txt": { "abstract": "Measuring g-mode pulsations of isolated white dwarfs can reveal their interior properties to high precision. With a spectroscopic mass of $\\approx 0.51 M_{\\odot}$ ($\\log g = 7.82$), the DAV white dwarf HS 1824+6000 is near the transition between carbon/oxygen core and helium core white dwarfs, motivating our photometric search for additional pulsations from the Palomar 60-inch telescope. We confirmed (with much greater precision) the three frequencies: $2.751190 \\pm 0.000010$ mHz (363.479 sec), $3.116709 \\pm 0.000006$ mHz (320.851 sec), $3.495113 \\pm 0.000009$ mHz (286.114 sec), previously found by B. Voss and collaborators, and found an additional pulsation at $4.443120 \\pm 0.000012$ mHz (225.067 sec). These observed frequencies are similar to those found in other ZZ Ceti white dwarfs of comparable mass (e.g. $\\log g<8$). We hope that future observations of much lower mass ZZ Ceti stars ($< 0.4 M_{\\odot}$) will reveal pulsational differences attributable to a hydrogen covered helium core. ", "introduction": "\\begin{deluxetable}{lcc} \\tablewidth{0pt} \\tablecaption{Properties of HS 1824+6000\\label{tab:prop}} \\tablehead{ \\colhead{ } & \\colhead{\\citealt{bv06}} & \\colhead{\\citealt{ag07}} } \\tablecolumns{3} \\startdata $T_{\\rm eff}$ (K) & $11192\\pm300$\\tablenotemark{a} & $11380\\pm140$\\tablenotemark{b} \\\\ $\\log g$ (dex) & $7.65\\pm0.10$\\tablenotemark{a} & $7.82\\pm0.04$\\tablenotemark{b} \\\\ $m_B$ (mag) & 15.7 & 15.7 \\\\ \\cutinhead{Observed Frequencies (in mHz)} \\citealt{bv06} & \\multicolumn{2}{l}{$2.6\\pm0.4$, $3.0\\pm0.9$, $3.3\\pm0.4$, $3.4\\pm0.8$} \\\\ This Paper & \\multicolumn{2}{l}{$2.751190 \\pm 0.000010$, $3.116709 \\pm 0.000006$} \\\\ & \\multicolumn{2}{l}{$3.495113 \\pm 0.000009$, $4.443120 \\pm 0.000012$} \\enddata \\tablenotetext{a}{Photometrically determined.} \\tablenotetext{b}{Spectroscopically determined.} \\end{deluxetable} \\begin{figure*}[t] \\centering \\epsscale{1.0} \\plotone{f1.eps} \\caption{The empirical ZZ Ceti instability strip. The circles represent systems for which temporal observations have been performed. Filled circles indicate systems not observed to vary while open circles indicate systems with observed periods. These data are from \\citet{pb04} and \\citet{ag05,ag07}. The vertical crosses represent low-mass WDs selected from the SDSS with $\\log g$ and $T_{\\rm eff}$ redetermined from MMT spectra from \\citet{mk07a}. The diagonal crosses are from \\citet{mk07a} except they are reanalysis of SDSS spectra and are merely candidate low-mass WDs until better spectra can be obtained. The square is LP 400-22 \\citep{ak06,mk07a}. The star is HS 1824+6000 (see Table \\ref{tab:prop}, \\citealt{ag07}). The error bar in the instability strip represents the typical error for measurements within the instability strip. The dashed lines are empirical fits to the instability strip as determined by \\citet{ag07}. The solid \\citep{jap07} and dash-dotted \\citep{lga01} lines correspond to cooling tracks of He WDs of the labeled mass. The labeled masses correspond to the following model masses (in $M_{\\odot}$) for \\citet{jap07} and \\citet{lga01} respectively: 0.19 to 0.1869 and 0.196, 0.24 to 0.2495 and 0.242, 0.40 to 0.3986 and 0.406, 0.45 to 0.4481 (\\citealt{jap07} only).} % \\label{fig:instrip} \\end{figure*} The hydrogen line ZZ Ceti variable (DAV) white dwarfs (WDs) occupy a discrete strip in the $T_{\\rm eff}-\\log g$ plane known as the ZZ Ceti instability strip. Many groups have assessed the location of this instability strip both empirically \\citep{fw91,asm04b,ag05,bgc07} and theoretically \\citep{pb97,yw99,gf03}, and despite minor discrepancies it spans $11,000 \\lesssim T_{\\rm eff} \\lesssim 12,250$ K for $\\log g \\approx 8.0$. \\citet{gf82} suggested that the instability strip is pure, meaning that all WDs within the strip are variable. However, \\citet{asm05} found numerous objects from the SDSS with associated low signal-to-noise spectra that were tenuously identified to be in the strip but did not vary to their observed detection limits. Recent observations \\citep{bgc07} have found some of these to be low amplitude pulsators. If the instability strip is pure, then it strongly implies that ZZ Ceti stars are a phase of evolution through which all DA WDs must evolve as they cool. White dwarfs less massive than $\\approx 0.45-0.47 M_{\\odot}$ ($\\log g \\approx 7.67$ at $T_{\\rm eff} \\approx 11,500$ K) \\citep{nld96,id99,ap04,dav06,jap07} do not undergo a He core flash in the course of their evolution and therefore are left with a He core. Two modes of evolution can truncate the red giant branch evolution and prevent the He core flash: mass loss due to winds and mass loss due to binary interaction. In systems of high metallicity, mass loss due to stellar winds on the red giant branch can be significant enough to lose the H envelope prior to the core flash \\citep{nld96,bmh05,mk07b}. Binary interaction through a common envelope also leads to significant mass loss \\citep{ii93,trm95}. Helium is thus the expected core composition for WDs below $\\approx 0.45-0.47 M_{\\odot}$. However, little direct evidence exists of the He core. Possible evidence would be the apparent over-brightness of old WDs \\citep{bmh05} in the star cluster NGC 6791 \\citep{lrb05}. Though uncertainties remain \\citep{cjd02,lrb08a,lrb08b}, recent detection of low $\\log g$ young WDs \\citep{jsk07} makes it plausible for many of the old WDs to be He core. A detailed asteroseismological study of these low-mass WDs could provide convincing evidence for the core composition. Observation and analysis of a full spectrum of the pulsation modes in a WD can produce a wealth of information about the interior structure of the WD. The mean period spacing of the modes, the rate of change in a mode's period over time, and multiplet splitting of individual modes can provide information on the total mass, spin rate, magnetic field strength, mass of H envelope, and core composition of the WD \\citep{ahc02,bgc08}. This has already been theoretically applied to distinguish between C/O and O/Ne core WDs by \\citet{ahc04}. The measured change in an observed mode period in G117-B15 has also been used to constrain significantly the C/O core composition of this object \\citep{sok91,sok95,sok00,sok05a}. We plot a version of the empirical instability strip in Figure \\ref{fig:instrip} . Included are not observed to vary (NOV) systems and pulsating ZZ Ceti stars from the observations of \\citet{pb04} and \\citet{ag05,ag07}. Also included are low-mass WDs from \\citet{mk07a}. There is a notable absence of low-mass ($\\log g \\lesssim 7.67$) WDs within the instability strip. There are a few possible ZZ Ceti stars of this low mass that are not plotted due to the absence of spectroscopically determined $\\log g$ and $T_{\\rm eff}$ measurements \\citep{bv07}. Also shown are the He WD cooling tracks of two models for $\\approx 0.19$, 0.24, 0.40, and $0.45 M_{\\odot}$ WDs \\citep{lga01,jap07}. The difference between these two models at low mass is due to the different H envelope masses. \\citet{lga01} used a $1 M_{\\odot}$ main sequence star and truncated its evolution up the red giant branch at various stages to produce He WDs of varying masses. \\citet{jap07} used close binary evolution expectations for main sequence stars of many masses to produce He WDs of varying masses. These different approaches cause the different remnant H envelope masses that yield a degeneracy in the He WD mass and its position in the $T_{\\rm eff}-\\log g$ plane. This degeneracy would be broken in the case of a ZZ Ceti He WD where the pulsation mode spectrum would reveal the H envelope mass. HS 1824+6000 (hereafter HS 1824, see Table \\ref{tab:prop}) was initially observed by \\citet{bv06} to exhibit pulsations. Their photometrically determined $\\log g$ and $T_{\\rm eff}$ placed its mass at $\\approx 0.40 M_{\\odot}$ using the tables of \\citet{lga97}. This mass was well within the theoretical expected mass range for He core WDs making it an excellent object to compare and contrast its pulsation frequencies with other C/O core and possible He core DAVs. However, later spectroscopic measurement by \\citet{ag07} determined its mass to be $\\approx 0.51 M_{\\odot}$, beyond the expected mass range for He core WDs. In \\S\\ref{sec:obsana} we discuss our own observations and differential photometry of HS 1824. In \\S\\ref{sec:timing} we apply a Lomb-Scargle Periodogram approach to a non-uniformly sampled time series in order to obtain the pulsation frequencies of HS 1824. In \\S\\ref{sec:flcpgs} we report the results of our observations and analysis. In \\S\\ref{sec:conc} we compare all observed ZZ Ceti periods with $\\log g < 8.0$. It is our hope this will yield a `zeroth-order' approach to He core identification in much the same way $T_{\\rm eff}$ and $\\log g$ measurements of field WDs identify likely ZZ Ceti stars. No singular distinction is currently present. \\clearpage ", "conclusions": "\\label{sec:conc} \\begin{figure} \\centering \\epsscale{1.0} \\plotone{f3.eps} \\caption{Top Panel: Lomb-Scargle Periodogram for combined data set (see \\S\\ref{sec:flcpgs}). Bottom Panel: Lomb-Scargle Periodogram for de-signaled (using only the four significant to 90\\% detected frequencies) data set . For both plots the dashed lines denote power level required for 90\\% significance.} \\label{fig:cspec} \\end{figure} \\begin{figure*} \\centering \\epsscale{1.0} \\plotone{f4.eps} \\caption{Spectrum of reported pulsation periods for published ZZ Ceti systems with spectroscopically determined $\\log g < 8$. HS 1824 is highlighted in gray with its four observed period locations marked with four vertical lines. Some marks represent more than one (very closely spaced) observed pulsation period, see references for details. HL Tau 76 lists only the verified independent pulsation modes of \\citet{nd06}. Several systems listed here are not included in Figure \\ref{fig:instrip} as their $\\log g$ and $T_{\\rm eff}$ measurements are not of sufficient precision. References: 1 - \\citealt{bgc06}, 2 - \\citealt{asm04a}, 3 - \\citealt{gv00}, 4 - \\citealt{rs05}, 5 - \\citealt{ag07}, 6 - \\citealt{ag06}, 7 - \\citealt{fm05}, 8 - \\citealt{pb04}, 9 - \\citealt{nd06}, 10 - \\citealt{pb95}, 11 - \\citealt{sok05b}, 12 - \\citealt{asm02}, 13 - \\citealt{asm03}.} \\label{fig:perspec} \\end{figure*} We have successfully detected four pulsation frequencies (periods), 2.751190 mHz (363.479 sec), 3.116709 mHz (320.851 sec), 3.495113 mHz (286.114 sec), and 4.443120 mHz (225.067 sec), in multiple observations of HS 1824+6000. There are also two possible pulsation frequencies (periods) at 4.450643 mHz (224.687 sec) and 5.755451 mHz (173.748 sec). With these periods of pulsation in HS 1824, the question remains if it, or other low gravity systems, can be empirically distinguished from the normal C/O core ZZ Ceti population. To answer this we compiled all known ZZ Ceti stars with published pulsation periods and spectroscopically measured gravities of $\\log g < 8.0$. This search resulted in 30 systems including HS 1824. In Figure \\ref{fig:perspec} we plot all reported periods for these 30 systems. Across all of these ZZ Ceti systems there exist many reported pulsation periods ranging from $100-1400$ sec. However, it is apparent that better than half of the reported periods reside within the range of $150-400$ sec. The four periods of HS 1824 are indistinguishable from the rest of this set of ZZ Ceti stars. Further, there does not appear to be any distinction between the two low-mass ($\\log g \\lesssim 7.67$) systems (HE 0031-5525, SDSS J2135-0743, \\citealt{bgc06}) and the rest of the set. With this current set of data it appears that this empirical analysis of reported pulsation periods is not sufficient to distinguish a suspected He core from a normal C/O core. However, HE 0031-5525, and SDSS J2135-0743 \\citep{bgc06} are very close to the boundary of He and C/O core WDs and within the errors of their $\\log g$ measurements may be C/O cores. It remains uncertain as to what degree this period spectrum comparative analysis can succeed. There are two primary differences between He and C/O core WDs that affect g-modes: the contrast in mean molecular weights in their cores, and the one fewer stratified layer in a He core object. G-modes penetrate deeply into the core, so that differences in the Brunt-V\\\"{a}is\\\"{a}l\\\"{a} profile there (due to the mean molecular weight; see \\citealt{cjd02}) significantly change the resulting mode period spectrum \\citep{pa06}. The stratified layers of material within the WD also affects how different pulsation modes are trapped, driven, and excited \\citep{ahc02,pa06}. He core WDs possess only two zones of He and H, while C/O core WDs possess the additional zone of C/O. Qualitatively, both of these differences would produce differences in the mode period spectra, and are the subject of current theoretical work we are pursuing. Once these full mode calculations are available, we can answer whether clear differences are observable. Most reported systems in Figure \\ref{fig:perspec} were found in observational campaigns looking only for pulsations in an effort to constrain the ZZ Ceti instability strip. In most cases, no attempt was made to distinguish observed pulsation periods as independent modes, as opposed to linear combinations of modes. This analysis was neglected in large part due to the lack of extensive follow up. Our observations of HS 1824 showed most single nights of data contain the pulsations of one specific period and it was a rarity to find a night of data with multiple pulsation periods. Ideally, very long gapless observations on the order of several days would address these problems very well. These observations could be obtained through the use of telescope networks such as the Whole Earth Telescope\\footnote{http://www.physics.udel.edu/darc/wet} as was done with HL Tau 76 \\citep{nd06} and G117-B15A \\citep{sok91,sok95} and the Las Cumbres Observatory Global Telescope\\footnote{http://www.lcogt.net}. We look toward future, more detailed observations of many low-mass and normal-mass ZZ Cetis to help provide a measurable distinction between He and C/O core compositions in WDs." }, "0808/0808.2210_arXiv.txt": { "abstract": "{We present the first part of a new catalog of variable stars (OIII-CVS) compiled from the data collected in the course of the third phase of the Optical Gravitational Lensing Experiment (OGLE-III). In this paper we describe the catalog of 3361 classical Cepheids detected in the $\\approx40$ square degrees area in the Large Magellanic Cloud. The sample consists of 1848 fundamental-mode (F), 1228 first-overtone (1O), 14 second-overtone (2O), 61 double-mode F/1O, 203 double-mode 1O/2O, 2 double-mode 1O/3O, and 5 triple-mode classical Cepheids. This sample is supplemented by the list of 23 ultra-low amplitude variable stars which may be Cepheids entering or exiting instability strip. The catalog data include {\\it VI} high-quality photometry collected since 2001, and for some stars supplemented by the OGLE-II photometry obtained between 1997 and 2000. We provide basic parameters of the stars: coordinates, periods, mean magnitudes, amplitudes and parameters of the Fourier light curve decompositions. Our sample of Cepheids is cross-identified with previously published catalogs of these variables in the LMC. Individual objects of particular interest are discussed, including single-mode second-overtone Cepheids, multiperiodic pulsators with unusual period ratios or Cepheids in eclipsing binary systems. We discuss the variations of the Fourier coefficients with periods and point out on the sharp feature for periods around 0.35~days of first-overtone Cepheids, which can be explained by the occurrence of 2:1 resonance between the first and fifth overtones. Similar behavior at $P\\approx3$~days for 1O Cepheids and $P\\approx10$~days for F Cepheids are also interpreted as an effect of resonances between two radial modes. We fit the period--luminosity relations to our sample of Cepheids and compare these functions with previous determinations.}{Cepheids -- Stars: oscillations -- Magellanic Clouds} ", "introduction": "The Optical Gravitational Lensing Experiment (OGLE) is a wide-field sky survey originally motivated by search for microlensing events (Paczy{\\'n}ski 1986). The observing strategy of the project is to regularly monitor brightness of about 200~million stars in the Magellanic Clouds and Galactic bulge in the time-scales of years. A~by-product of these observations is an enormous database of photometric measurements, which can be used for selecting long lists of newly discovered variable stars. The OGLE project yielded a wealth of information about variable stars. The second phase of the survey, conducted between 1997 and 2000, resulted in catalogs of thousands Cepheids, RR~Lyr stars, eclipsing binaries and long period variables in the Magellanic Clouds. Moreover, the huge catalogs of variable sources found in the OGLE-II fields in the Magellanic Clouds ({\\.Z}ebru{\\'n} \\etal 2001) and in the Galactic bulge (Wo{\\'z}niak \\etal 2002) were released. In this paper we present the first part of the OGLE-III Catalog of Variable Stars (OIII-CVS) -- the catalog containing virtually all variable stars in the fields regularly observed by OGLE since 2001 with the Warsaw telescope at Las Campanas Observatory, Chile. Classical Cepheids ($\\delta$~Cep stars, type I Cepheids), as the primary distance indicator, are among the most important variable stars. The Large Magellanic Cloud (LMC) is one of the most fundamental extragalactic targets of modern astrophysics, because it is our nearest non-dwarf neighbor galaxy. For this reason we begin the OIII-CVS with the catalog of classical Cepheids in the LMC. Large number of variable stars in the LMC, including classical Cepheids, were discovered by Leavitt (1908). However, the first period derivation and the plot of the period--luminosity (PL) diagram for 40 LMC Cepheids was made by Shapley (1931). In 1955 the periods of 550 classical Cepheids in the LMC were published (see Shapley and McKibben Nail 1955 for the bibliography). Then, a considerable survey for LMC Cepheids was done by Woolley \\etal (1962). The catalog prepared by Payne-Gaposchkin (1971) on the basis of Harvard photographic plates contained about 1100 Cepheids in the LMC. After hiatus, a number of Cepheids in the LMC were also discovered by Hodge and Lee (1984), Kurochkin \\etal (1989), van Genderen and Hadiyanto Nitihardjo (1989) and Mateo \\etal (1990). In the late 1990's very large catalogs of Cepheids were published as a by-product of gravitational microlensing surveys: EROS (Beaulieu \\etal 1995, Afonso \\etal 1999), MACHO (Welch \\etal 1997, Alcock \\etal 1999b) and OGLE-II (Udalski \\etal 1999d, Soszy{\\'n}ski \\etal 2000). The catalog described in this work contains the largest sample of classical Cepheids detected to date in the LMC and, likely, in any other environment. Almost 1000 objects are new identifications. Double-mode Cepheid sample presented in this paper is three times more numerous than the largest sample presented so far. We also show individual objects of particular interest, like triple-mode Cepheids, Cepheids with non-radial pulsations, Cepheids in eclipsing binary systems, etc. The paper is organized as follows. In Section~2 we describe how the observations were obtained and reduced. Section~3 gives the details about the process of Cepheid selection. In Section~4 we describe the catalog itself. We compare our sample with previously published catalogs of the LMC classical Cepheids in Section~5. In Section~6 we discuss the Fourier coefficients as a function of periods. In Section~7 we fit the period--luminosity relations. Finally, Section~8 summarizes the paper. ", "conclusions": "In this paper we present the largest catalog of classical Cepheids in the LMC and probably the largest sample of such stars identified to date in any environment. Our list of Cepheids is supplemented by the high quality, long-term standard photometry enabling precise analysis of these stars. These data are ideal for studying many fundamental problems, such as interpretation of the pulsational and evolutionary models of Cepheids, non-radial oscillations in the pulsating stars, possible non-linearity of the PL relation, structure and history of the LMC. The catalog contains very rare objects, such as Cepheids with three radial modes excited, 1O/3O double-mode Cepheids, single-mode second-overtone pulsators, Blazhko Cepheids, eclipsing binary systems containing Cepheids including system of two Cepheids eclipsing each other. Our data show that first-overtone classical Cepheids and high amplitude $\\delta$~Sct stars follow continuous PL relation. Distribution of the Fourier parameters suggests that the internal resonance between radial modes may occur twice for the first-overtone pulsators: for periods of about 0.35~days and 3~days. The PL relation for first-overtone Cepheids is possibly non-linear, with a~discontinuity in the slope around $P=0.5$~days. In the next parts of the OIII-CVS we will present other members of the Cepheid family: type II Cepheids, anomalous Cepheids, HADS and RR~Lyr stars. It is possible that the list of classical Cepheids described in this paper will be supplemented with additional objects of this type detected during further analysis. \\Acknow{The authors wish to thank Prof.~W.A.~Dziembowski, Prof.~M.~Feast, and Prof.~W.~Gieren for many helpful suggestions which improved the paper. We thank Drs. Z.~Ko{\\l}aczkowski, T.~Mizerski, G.~Pojma{\\'n}ski, A.~Schwar\\-zenberg-Czerny and J.~Skowron for providing the software and data which enabled us to prepare this study. This work has been supported by the Foundation for Polish Science through the Homing (Powroty) Program and by MNiSW grants: NN203293533 to IS and N20303032/4275 to AU. The massive period searching was performed at the Interdisciplinary Centre for Mathematical and Computational Modeling of Warsaw University (ICM UW). We are grateful to Dr. M.~Cytowski for helping us in this analysis.}" }, "0808/0808.2483_arXiv.txt": { "abstract": "{Blazars are thought to emit highly-collimated outflows, so-called jets. By their close alignment to our line of sight, relativistic beaming effects enable us to observe these jets over the whole electromagnetic spectrum up to TeV energies, making them ideal laboratories for studying jet physics. In the last years multiwavelength observations of blazars provided us with detailed data sets which helped to characterize the two main components of the non-relativistic emission, peaking in the optical to X-ray and GeV/TeV energy region, respectively. In leptonic acceleration models, they are explained by synchrotron radiation of electrons and inverse-Compton emission from the same electron population and thus, correlations of both emission regimes are expected. We review recent observational results on the presence and absence of such correlations in blazars, and discuss constraints on emission models by quantitative correlation analyses.} \\FullConference{Workshop on Blazar Variability across the Electromagnetic Spectrum\\\\ April 22-25 2008\\\\ Palaiseau, France} \\begin{document} ", "introduction": " ", "conclusions": "The up to now best-studied objects concerning correlations of the two bumps in the spectral energy distribution of blazars all belong to the class of high-peaked BL~Lac objects. These bright TeV blazars have been studied in great detail for about the last ten years on various timescales and during various emission-level episodes. Correlated variability up to now has been observed on timescales ranging from hours to months, up to years and seems to be an observationally established fact for the well-studied TeV blazars. Most of the results, however, still suffer from various possible experimental caveats, including the need for an as-simultaneous-as-possible time coverage and a good sensitivity of the instruments included in the observational campaigns. Last but not least, one needs some luck to be able to follow a blazar during an outburst and be able to infer the physics of the flare production and decay. There have been, however, few campaigns, as e.g. the 2001 campaign on Mkn 421, which provided the possibility to study correlations during the rise and decay phase of individual flares; particularly also the observation of the 2006 flares of PKS 2155--304 seem to offer almost a dissecting knife for studying the quantitative imprints of the physics processes that produce flares on the correlations. At the same time, interesting phenomena like ``orphan flares'' without X-ray counterparts and ``childless X-ray flares'' need to be explained, and more generally, a systematic understanding of lags between the two photon populations is still elusive. While correlations of the emission on the high-energy tails of the two photon populations have been studied in great detail, lately also potential correlations of VHE $\\gamma$ rays with the optical or radio emission have been investigated. Optical triggers seem to be a successful proxy to finding new TeV emitters, while first studies of radio-TeV correlations seem to help locating the region in which the SSC phenomenon takes place, profiting from the high spatial resolution that VLBI can offer." }, "0808/0808.0165_arXiv.txt": { "abstract": "The CoNFIG (Combined NVSS-FIRST Galaxies) sample is a new sample of 274 bright radio sources at 1.4~GHz. It was defined by selecting all sources with \\SoneG$\\ge$1.3~Jy from the NRAO-VLA Sky Survey (NVSS) in the North field of the Faint Images of the Radio Sky at Twenty-cm (FIRST) survey. New radio observations obtained with the VLA for 31 of the sources are presented. The sample has complete FRI/FRII morphology identification; optical identifications and redshifts are available for 80\\% and 89\\% of the sample respectively, yielding a mean redshift of $\\sim$0.71. One of the goals of this survey is to get better definitions of luminosity distributions and source counts of FRI/FRII sources separately, in order to determine the evolution of the luminosity function for each type of source. We present a preliminary analysis, showing that these data are an important step towards examining various evolutionary schemes for these objects and to confirm or correct the dual population unified scheme for radio AGN. Improving our understanding of radio galaxy evolution will give better insight into the role of AGN feedback in galaxy formation. ", "introduction": "\\cite{Long66} determined that powerful radio sources undergo strong differential evolution, the first indication of cosmic downsizing. Since then our understanding of the space density of AGN as a function of cosmic epoch has steadily continued to advance.\\\\ \\indent With the development of evolutionary models for radio sources came the idea of a dual-population model. The initial version of this dichotomy \\citep{Long66} was in terms of low and high luminosities. But with high-frequency surveys, and the large number of flat/inverted spectrum sources revealed in them, an alternative classification emerged, based exclusively on the source spectra: sources with a spectral index $\\alpha \\le -0.5$ (where $S^{\\alpha}_{\\nu} \\propto \\nu^{\\alpha}$), corresponding to optically thin synchrotron radiation, were classified as steep-spectrum, whereas sources with lower spectral indices ($\\alpha \\ge -0.5$) inevitably showed features of synchrotron self-absorption and were classified as flat-spectrum. Initial indications suggested that these two populations underwent different evolution \\citep{Schmidt76,Masson77}. However, \\cite{Dunlop90} studied the radio luminosity function (density of sources with a given luminosity per unit of co-moving volume) of these two classes of radio sources, and came to the conclusion that both populations were undergoing a similar evolution, implying that they might not actually be distinct.\\\\ \\indent The Fanaroff-Riley (FR) classification \\citep{FR74}, originally a sub-classification for steep-spectrum objects, was employed to provide a further categorization of radio sources. This classification divides radio sources into two classes of double-lobed sources based on the appearance of their jets. The FRI objects have the highest brightness along the jets and core, reside in moderately rich cluster environments \\citep{Hill91} and include sources with irregular structure \\citep{Parma92}. In contrast, FRII sources show hot spots in the lobes and more collimated jets, are found in more isolated environments and display stronger emission lines \\citep{Rawlings89,Baum89}. \\cite{FR74} found these two classes to be divided in radio power, with a break luminosity $P_{\\rm 178~MHz} \\sim 10^{25}\\, {\\rm W Hz^{-1}sr^{-1}}$, with FRII sources lying above this limit. Subsequently \\cite{Owen94} showed that the break was a function of both radio and optical luminosity.\\\\ \\indent During the 1980s the 'unification' hypothesis emerged to describe how viewing aspect could relate RQSOs, (Radio Quasi-Stellar Objects of either flat or steep-spectrum) to FRII radio galaxies \\citep[e.g.][]{Peacock87,Scheuer87,Barthel89}. However, the scheme did not include lower-luminosity AGNs such as FRI galaxies and BL Lac objects. The unifying connection between these was introduced by \\cite{Marcha95}. The unified model of AGN proposed by \\cite{Wall97} and \\cite{Jack99} assumes that the cosmic evolution of radio loud AGN is based on a division of the radio sources into a low-luminosity ($P_{\\rm 178~MHz}<10^{25}\\, {\\rm W Hz^{-1}sr^{-1}}$) component corresponding to FRIs, and a high-luminosity component corresponding to FRIIs. In this scheme, the various forms of AGN observed (FRI and FRII extended double sources, flat- and steep-spectrum RQSOs and BL~Lac objects) result from the orientation of the extended parent objects with respect to the observer's line-of-sight. Indeed, because the double-sided ejection of synchrotron blobs in AGN is at relativistic speed, the orientation of the ejection axis to the line-of-sight becomes crucial: sources viewed side-on appear as double radio galaxies (FRI or FRII) and sources viewed along the jets appear as RQSOs (beamed counterparts of FRII sources) or BL~Lac objects (beamed counterparts of FRI sources). The relativistically-boosted jet emission in the beamed counterparts of the extended sources dominates the extended emission, making the overall radio emission appear compact down to VLBI scales.\\\\ \\indent Initially, in modelling the space density of radio AGN, it was assumed for simplicity that the low-luminosity radio galaxies including FRI sources showed no cosmic evolution \\citep{WPL80,Jack99}, the strong cosmic evolution confined only to the higher luminosities and the FRII galaxies. With the advent of large-scale redshift surveys for nearby galaxies, many authors, including \\cite{Brown01}, \\cite{Snellen01}, \\cite{Willott01}, \\cite{Sadler07} and \\cite{Rigby07}, found significant evolution for low power sources -- but mild evolution in comparison with that of the high-luminosity sources. \\cite{Rigby07} argued that if both FRIs and FRIIs have similar evolution, the dual-population scheme could be reduced to a single-population model.\\\\ The FRI/FRII dual-population scheme has encountered several problems. One of these concerns the correspondence between FRI galaxies and BL~Lac objects. \\cite{Urry95} noted that some BL~Lac objects have non-FRI-like morphologies and that the density of FRI sources is too low to account for the entire BL~Lac population, a concern also raised by \\cite{Wall97}. Looking at BL Lac objects from another point of view, \\cite{March96} demonstrate that only about one third of low-luminosity core dominated radio sources - which are supposedly the beamed counterpart of FRI sources - are conventional BL~Lac objects. Most of the remaining sources have optical classification such as Seyfert objects or elliptical galaxies.\\\\ \\indent A related issue concerns the existence of FRI RQSOs. Until recently, these objects were thought not to exist, leading to the hypothesis that FRI and FRII central engines were of different nature \\citep{Baum95} and that the torus opening in FRI sources was too small to observe a quasar nucleus \\citep{Falcke95}. However, the discovery of an FRI QSO, E1821+643, by \\cite{Blun01} overthrew those assumptions. More recent VLA observations \\citep{Heywood07} uncovered another 4 sources of this type.\\\\ \\indent Finally, if sources with different FR classes undergo different evolution, this might imply that their fundamental characteristics, such as the black hole spin or jet composition, are different too. However, the existence of hybrid sources, which display both FRI and FRII morphological characteristics \\citep{Capetti95}, then becomes puzzling. Based on observations of hybrid sources, \\cite{Kaiser97} argued that the FR dichotomy is based purely on the interaction between the jets and the environmental medium, and not on intrinsic properties of the central engine. This view is also supported by \\cite{Gopal00} and \\cite{Gawr06}. However, \\cite{Wang92} suggested that some AGN engines could be capable of ejecting jets of unequal power, resolving the problem of hybrid sources. \\cite{Gopal00} found no evidence for such a process in their sample.\\\\ \\indent Other schemes have been suggested to resolve these difficulties. \\cite{Kaiser07} explained the FR dichotomy by postulating that all sources start as FRII objects and used an analytical model in which the evolution of the radio sources is governed by energy losses from both radiating relativistic electrons in the lobes and turbulence in the jets.\\\\ \\indent \\cite{Willott01} used an approach to a dual-population unified scheme based on the luminosity of sources instead of morphology. Optical spectra of FRII sources are heterogeneous and they can display both strong and weak low-excitation emission lines \\citep{Laing94}. Therefore, radio sources can also be grouped based on their emission lines, with one population composed of low-luminosity sources having weak emission lines (containing both FRI and FRII objects), and the other composed of high-luminosity sources with strong emission lines (containing only FRII objects). With this model, \\cite{Willott01} concluded that high luminosity objects undergo a stronger evolution with epoch than low luminosity sources, as found by by e.g. \\cite{Long66}, \\cite{WPL80} and \\cite{Urry95}, and that the radio luminosity function has the form of a broken power-law, similar to the conclusions of \\cite{Dunlop90}. We note that, since conventional accretion-disk systems are expected to be strong X-ray sources, luminosity-based dual-population models are also used in modelling the luminosity function of X-ray selected RQSOs \\citep{Hardcastle06}.\\\\ Defining the relation between the different radio sub-populations together with their cosmic evolution is becoming fundamental to our understanding of galaxy formation.\\\\ The current paradigm for galaxy formation follows hierarchical build-up in a Cold Dark Matter (CDM) universe. Nevertheless, serious difficulties arise from this model in its simplest form, as discussed by \\cite{Bower06}. It implies that current epoch galaxies must be the largest and bluest and have the highest star forming rate of all galaxies. Observations show that they are red, old galaxies, whereas the bulk of star formation is observed at earlier epochs. This is known as downsizing, first described by \\cite{Cowie96}. AGN negative feedback could be the key to understanding this phenomenon. The ignition of the nucleus in a star forming galaxy could eject the gas into the inter-galactic medium, thus reducing or even stopping star formation, breaking the hierarchical buildup \\citep{Silk98,Granato01,Quilis01}. Note that there is also some positive AGN feedback \\citep{vanB04,Klamer04} in which the pressure from the jets compresses the inter-stellar medium and induces star formation. The balance between these processes remains to be understood; establishing the cosmic behaviour of the radio AGN is important in revealing the precise role of the feedback mechanisms.\\\\ Although FRI and FRII sources show different evolution, they also lie in different luminosity ranges. There is therefore a possibility that both types may show similar evolution for overlapping luminosities (i.e. high-luminosity FRIs and low-luminosity FRIIs). In order to sort out the FR dichotomy and its details, accurate models of the evolution of each population are needed. This implies compiling accurate statistics, such as luminosity distributions and source counts, for both types separately. This is the goal of the CoNFIG (Combined NVSS-FIRST Galaxies) sample presented here, a new sample of bright radio sources with complete morphological identification.\\\\ \\indent The CoNFIG sample is defined as all sources with \\SoneG$\\ge$1.3~Jy from the NRAO-VLA Sky Survey \\citep[NVSS,][]{Condon98} catalogue within the north region of the Faint Images of Radio Sources at Twenty-cm survey \\citep[FIRST,][]{White97}, a 1.5~sr region defined roughly by $-8^{\\circ} \\le {\\rm Dec}\\le 64^{\\circ}$ and 7 hr $\\le {\\rm RA}\\le$ 17 hr. The flux density limit of 1.3~Jy was chosen so that the number of sources in the sample was of statistical significance while allowing us to identify the morphology for each source individually. Optical identifications were obtained from the SuperCOSMOS Sky Survey \\citep[SSS,][]{Hambly01} and redshift information extracted using the SIMBAD database. With the accompanying VLA observations described here, the sample has complete morphology information, a median flux density of \\SoneG$\\sim$1.96~Jy and optical identifications and redshift information for $\\sim$80\\% and $\\sim$89\\% of the sources respectively.\\\\ The structure of this paper is as follows. The details of the construction of the CoNFIG sample are explained in \\S\\ref{sample} while \\S\\ref{morpho} describes how the morphologies were determined. Optical identifications and redshift information are discussed in \\S\\ref{IDandz} and \\S\\ref{lumdist} outlines the computation of the morphology-dependent luminosity distributions. Finally, \\S\\ref{extSC} describes and discusses the FRI/FRII source counts.\\\\ \\indent Throughout this paper, we assume a standard $\\Lambda$CDM cosmology with \\Hzero=70 km s$^{-1}$ Mpc$^{-1}$, \\OM=0.3 and \\OL=0.7. ", "conclusions": "The CoNFIG sample is constructed as a sample of 274 radio sources from NVSS with $S\\ge 1.3$~Jy. Redshift information is available for $\\sim$80\\% of the sample, and morphological classifications were obtained for all of the sources, either from NVSS and FIRST contour plots, or, for 46 sources, from 8~GHz VLA observations. These data allow us to compute morphology-dependent luminosity distributions and source counts.\\\\ \\indent To increase the number of sources with morphology information, three more samples were constructed in sub-areas of the main region with flux density limits of 0.8~Jy, 0.2~Jy and 50~mJy. Morphological identifications were obtained only from NVSS and FIRST contour plots for those sources. Morphological information for the CENSORS and BDFL sample were obtained from the Ledlow-Owen relation between radio power and optical magnitude and from published classification respectively. Combining these six samples allowed us to compile source counts for FRI and FRII sources separately. A simple, single evolution model for space density was then fitted to these data.\\\\ \\indent Our data show mild evolution of the FRI sources at low redshift; however, they do not participate in the ``evolution bump'' around \\SoneG$\\sim$1~Jy. The results also support the observation that a large number of mJy sources are FRIs galaxies and not starburst galaxies as previously assumed.\\\\" }, "0808/0808.0171_arXiv.txt": { "abstract": "We present results of VLBI observations of the water masers associated with IRAS 4A and IRAS 4B in the NGC 1333 star-forming region taken in four epochs over a two month period. Both objects have been classified as extremely young sources and each source is known to be a multiple system. Using the Very Long Baseline Array, we detected 35 masers in Epoch I, 40 masers in Epoch II, 35 in Epoch III, and 24 in Epoch IV. Only one identified source in each system associates with these masers. These data are used to calculate proper motions for the masers and trace the jet outflows within 100 AU of IRAS 4A2 and IRAS 4BW. In IRAS 4A2 there are two groups of masers, one near the systemic cloud velocity and one red-shifted. They expand linearly away from each other at velocities of 53 \\kms. In IRAS 4BW, masers are observed in two groups that are blue-shifted and red-shifted relative to the cloud velocity. They form complex linear structures with a thickness of 3 mas (1 AU at a distance of 320 pc) that expand linearly away from each other at velocities of 78 \\kms. Neither of the jet outflows traced by the maser groups align with the larger scale outflows. We suggest the presence of unresolved companions to both IRAS 4A2 and 4BW. ", "introduction": "Water masers are excellent probes of astrophysical flows. Their large flux densities and exceedingly compact sizes make them perfect Very Long Baseline Interferometer (VLBI) targets. Many masers observed in star-forming regions are thought to form in shocks either in or along the outflow of material commonly seen associated with young stellar objects (YSOs). They provide a high resolution probe of the base of the stellar wind that is unaffected by extinction from dust or extensive interaction with ambient material. The milliarcsecond resolution of VLBI is critical to resolve the individual maser components and outflows, particularly in high stellar density environments. Long term total power monitoring has shown that masers around low mass YSOs are more episodic than those around higher mass YSOs, however, their detectable phases are sufficiently long for proper motion measurements if the observations are spaced by no more than two to three weeks \\citep{Furuya:2003,Brand:2003, CLAUSSEN:1996,WILKING:1994}. Furthermore, if they arise in the warm (400K) spatially confined post-shock region, as postulated in various models \\citep{MODEL1, MODEL2, MODEL3, MODEL4}, they should have space velocities sufficient to produce measurable proper motions over one to three weeks. However, regular monitoring of low mass YSOs is necessary to ensure that VLBI observations are conducted during periods of high maser activity. Using the NRAO's \\footnote{The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.} Very Long Baseline Array (VLBA), proper motion studies have been made of water masers associated with high mass and intermediate mass YSOs such as Cepheus A HW2 \\citep{TORRELLES:2001}, S~106 FIR \\citep{FURUYA:2000}, NGC 2071 \\citep{SETH:2002}, and IRAS~20050+2720 MMS1 \\citep{FURUYA:2005} as well as with lower mass YSOs such as Serpens SMM1, RNO 15-FIR, and IRAS~05413-0104 \\citep{Moscadelli:2006,CLAUSSEN:1998}. For example, water masers in IRAS~05413-0104 (associated with HH~212 at a distance of 450 pc), were found to lie in a structure of 10 AU in length and less than 0.5 AU thick. The masers appeared to arise in shock-related structures which showed proper motions along the axis of the outflow of 60 \\kms and displayed coherent structures over time scales of two to three weeks (but which was less discernible over time scales as long as several months). Comparison with molecular and infrared observations of IRAS~05413-0104 clearly demonstrated that proper motion studies of water masers in the region (within 40 AU) of the central source are among the best tools for studying the kinematics of the jets emanating from embedded YSOs. IRAS~4, comprised of multiple sources whose submillimeter-dominated 'Class 0' spectral energy distributions suggest extreme youth, is located in the NGC 1333 star-forming region. As a nearby active and multiple young object, it presents a prime candidate for VLBI observations of its associated water masers. Located at a distance of 320 pc \\citep{HIPPARCOS}, the NGC 1333 molecular cloud hosts a double infrared cluster of about 200 YSOs identified in near-infrared surveys \\citep{STROM:1976,ASPIN:1994,Lada:1996,Wilking:2004} and x-ray surveys \\citep{PREIBISCH:1997,Getman:2002,PREIBISCH:2003}. Far-infrared observations have not only revealed the higher luminosity sources in NGC 1333, but also YSOs in the earliest phase of evolution such as the Class 0 sources IRAS~4 and IRAS~2 \\citep{Harvey:1984,JENNINGS:1987}. IRAS~4 was found to be a binary at submillimeter wavelengths \\citep{SANDELL:1991}; hence IRAS~4A and 4B. IRAS~4A is located 31$\\arcsec$ to the NW of 4B at a position angle of -45$^{\\circ}$. Subsequently, both IRAS~4A and 4B were found to be binary systems. \\citet{LAY:1995} performed single baseline interferometry in the submillimeter and found that 4A was a binary with an angular separation of 1.8$\\arcsec$ (or 580 AU at 320 pc). Interferometric observations in the millimeter and radio continuum confirmed the separation and determined that the position angle of separation was $\\approx -20^{\\circ}$ \\citep{LOONEY:2000,REIPURTH:2002}. While the easternmost source (IRAS~4A1) dominates the millimeter and radio continuum emission, the westernmost source IRAS~4A2) appears to be relatively more evolved with a warm ammonia core \\citep{WOOTTEN:1993,Shah:2000} and a wealth of complex organic molecules (Bottinellii 2007, priv. comm.). Submillimeter and millimeter wave continuum observations have confirmed the prediction of \\citet{LAY:1995} that IRAS~4B is also a binary with components (4BE and 4BW) separated by about 10$\\arcsec$ (3200 AU in projection) along an E-W direction \\citep{SMITH:2000,LOONEY:2000,Sandell:2001}. \\footnote{We adopt the naming convention introduced by \\citet{Sandell:2001}. The western (eastern) source is referred to as IRAS~4BI (IRAS~4BII) by \\citet{SMITH:2000}, as IRAS~4B (IRAS~4C) by \\citet{LOONEY:2000}, and as IRAS~4B (4B$\\arcmin$) by \\citep{DIFRANCESCO:2001}.} As in the case of IRAS~4A2, complex organic molecules have been detected toward IRAS~4B \\citep{Bottinelli:2007}, presumably from a hot core associated with 4BW. We note that IRAS~4BE has yet to be detected in the radio continuum \\citep{REIPURTH:2002} or demonstrate any sign of outflow activity. Infall motions have been detected toward IRAS~4A and 4B \\citep{DIFRANCESCO:2001}, unambiguously toward the former, and strongly indicated toward the latter. The detection of water masers toward IRAS~4A and 4B suggests that they are YSOs also associated with mass outflow. Water masers were first detected in single dish observations by \\citet{HASCHICK:1980} and later monitored by \\citet{CLAUSSEN:1996}. These observations established the maser emission as highly variable on monthly time scales, sometimes being completely absent while at other times reaching peaks of 10 Jy. Detected emission was found from $-$10 to $+$15 \\kms but the angular resolution was not sufficient to determine which masers were associated with 4A and 4B. Beginning in 1983, VLA observations have shown water maser activity associated only with IRAS~4A2 and/or 4BW \\citep{RODRIGUEZ:2002,Wootten:1998,Furuya:2003}. VLBA observations of the water masers in IRAS~4A and 4B, acquired in 2003, have been presented by \\citet{Desmurs:2006}. While masers were detected in only two of the four epochs observed, they confirmed the association of masers with IRAS~4A2 and expansion motion between the two spots was detected. In IRAS~4BW, the five maser spots detected were red-shifted and formed a chain 80 mas in extent with some masers displaying proper motion toward the north. Molecular outflows have been mapped toward both IRAS~4A and 4B. IRAS~4A is associated with a highly collimated outflow about 20,000 AU in extent seen in HCN and SiO with a position angle close to 20$^{\\circ}$ \\citep{Girart:1999,Choi:2001,Choi:2005}. The origin of the outflow is likely IRAS~4A2 \\citep{Choi:2005}. On a larger scale ($\\approx 1 \\times 10^{5}$ AU), the outflow defined by CO and molecular hydrogen has a position angle of 45$^{\\circ}$ \\citep{BLAKE:1995,Choi:2006}. The shift in position angle from small to large scale is perhaps due to a combination of a shift in the direction of the magnetic field, an encounter with denser ambient gas, and precession of the outflow axis \\citep{Choi:2006}. A more compact outflow is associated with IRAS~4BW, with a position angle close to 0$^{\\circ}$ \\citep{BLAKE:1995,Choi:2001} with perhaps a second outflow with a position angle of -35$^{\\circ}$ \\citep{DIFRANCESCO:2001}. In this paper we present VLBA observations of water masers associated with the IRAS~4 region obtained over four epochs in 1998 and spaced by about one month. We also describe Very Large Array (VLA) observations, taken one month in advance of the start of the VLBA observations, that showed IRAS~4 to be a very active maser source and yielded absolute positions for the masers and the 1.3 cm continuum sources IRAS~4A1 and IRAS~4BW. With the VLBA, we detected 35 masers in Epoch I, 40 masers in Epoch II, 35 in Epoch III, and 24 in Epoch IV associated with IRAS~4A2 and IRAS~4BW. We use these data to reveal the structure of shock fronts associated with stellar winds from these YSOs and to estimate proper motions for masers detected in all four epochs. The origin of the maser emission and its relationship to the stellar winds and larger scale molecular outflows from these YSOs is discussed. ", "conclusions": "We have observed the water masers associated with IRAS~4A and 4B at VLBI resolutions in four epochs over three months. We have determined that the masers are related to the jets emanating from these YSOs due to their spatio-kinematic distribution. In both sources the masers are found associated with known single components of multiple systems, with total separation velocities between 53 and 78 km s$^{-1}$. This is further confirmed by the large proper motions measured for both sources, which clearly rules out rotation due to the mass constraints placed on the central objects by other observations. The water masers of IRAS~4B form arc-like structures roughly 10 AU in length and less than 0.6 AU in thickness. The structure of these structures changes rapidly with time, with many new maser components appearing and disappearing in just one month. The orientation of the structures in the plane of the sky does not agree with larger scale outflow angle, which we attribute to a possible unseen very close companion. Future observations will have to sample the source more frequently than once every three weeks, with once every 3-5 days probably providing the best results." }, "0808/0808.2497_arXiv.txt": { "abstract": "Many current and future astronomical surveys will rely on samples of strong gravitational lens systems to draw conclusions about galaxy mass distributions. We use a new strong lensing pipeline (presented in Paper~I of this series) to explore selection biases that may cause the population of strong lensing systems to differ from the general galaxy population. Our focus is on point-source lensing by early-type galaxies with two mass components (stellar and dark matter) that have a variety of density profiles and shapes motivated by observational and theoretical studies of galaxy properties. We seek not only to quantify but also to understand the physics behind selection biases related to: galaxy mass, orientation and shape; dark matter profile parameters such as inner slope and concentration; and adiabatic contraction. We study how all of these properties affect the lensing Einstein radius, total cross-section, quad/double ratio, and image separation distribution, with a flexible treatment of magnification bias to mimic different survey strategies. We present our results for two families of density profiles: cusped and deprojected S\\'ersic models. While we use fixed lens and source redshifts for most of the analysis, we show that the results are applicable to other redshift combinations, and we also explore the physics of how our results change for very different redshifts. We find significant (factors of several) selection biases with mass; orientation, for a given galaxy shape at fixed mass; cusped dark matter profile inner slope and concentration; concentration of the stellar and dark matter deprojected S\\'ersic models. Interestingly, the intrinsic shape of a galaxy does not strongly influence its lensing cross-section when we average over viewing angles. Our results are an important first step towards understanding how strong lens systems relate to the general galaxy population. ", "introduction": "\\label{S:motivation} The field of strong lensing is undergoing a period of growth that is expected to accelerate in the coming decade. Current samples with tens of lenses, such as from CASTLES\\footnote{CASTLES is a collection of uniform HST observations of mostly point-source lenses from several samples with differing selection criteria, rather than a single, uniformly-selected survey.} \\citep[e.g.,][]{2001ASPC..237...25F}, CLASS \\citep[e.g.,][]{2003MNRAS.341...13B}, SDSS \\citep[e.g.,][]{2008AJ....135..496I}, and SLACS \\citep[e.g.,][]{2008ApJ...682..964B}, will give way to samples with hundreds or even thousands of strong lensing systems discovered by Pan-STARRS, LSST, SNAP, and SKA \\citep[e.g.,][]{ 2004AAS...20510827F, 2004NewAR..48.1085K, 2004ApJ...601..104K, 2005NewAR..49..387M}. As lens samples grow, they will be even more useful for constraining the mass distributions of the strong lensing galaxy population \\citep[for a review of strong lensing astrophysics, see][]{Saas-Fee}. There are, however, certain complications in using strong lensing studies to constrain the physical properties of galaxies. The issue we address here is \\emph{selection bias}. Suppose we treat galaxies as having an intrinsic probability distribution in some parameter $x$ (such as the inner slope of the density profile), and we want to use observed strong lens systems to infer that distribution $p(x)$. If the lensing probability itself depends on $x$, then in general the distribution $p_\\mathrm{SL}(x)$ for the strong lens galaxies will not reflect the true, underlying $p(x)$ for all galaxies. The more diverse the galaxy population, and the more the strong lensing cross-section varies across the population, the more important the selection biases can be. Obviously, one must account for selection biases in order to draw reliable conclusions from current and future strong lens samples. Our purpose here is to use a complete, end-to-end, and above all, realistic simulation pipeline for strong lensing (presented in Paper~I of this series, van de Ven, Mandelbaum, \\& Keeton 2008) to quantify and understand strong lensing selection biases and assess their impact in typical situations. We focus on strong lensing of quasars by early-type, central (non-satellite) galaxies at several mass scales, using two-component mass profiles (dark matter plus stars) that are consistent with existing photometry and stacked weak lensing data from SDSS \\citep{2003MNRAS.341...33K,2006MNRAS.368..715M} as well as with $N$-body and hydrodynamic simulations. The issue of selection biases in strong lensing is not a new one, and we are building on many previous studies of this issue. The main galaxy properties that we study in an attempt to understand selection bias are \\begin{itemize} \\item Mass \\citep[e.g., ][]{1984ApJ...284....1T, 1991MNRAS.253...99F, 2007MNRAS.379.1195M}; \\item Orientation \\citep[``inclination bias'', e.g., ][]{1997ApJ...486..681M, 1998ApJ...495..157K, 2007arXiv0710.1683R}; \\item Shape \\citep[e.g., ][]{1997ApJ...482..604K, 2005ApJ...624...34H, 2007arXiv0710.1683R}; \\item Dark matter inner slope \\citep[e.g., ][]{2001ApJ...549L..25K, 2001ApJ...555..504W,2002ApJ...566..652L}; \\item Dark matter concentration \\citep[e.g., ][]{2004ApJ...601..104K, 2007A&A...473..715F}. \\end{itemize} In most cases, previous studies addressed the issue of selection bias using models that were simplified in some way. Simplifications often included testing effects of dark matter slope and concentration using pure (generalised) NFW profiles without a baryonic component; or testing effects of density profile or shape using single-component mass models. A notably different approach was taken by \\citet{2007MNRAS.382..121H}, \\citet{2007MNRAS.379.1195M}, and \\citet{2008MNRAS.386.1845H}, who used the Millennium simulation with a semi-analytic model of galaxy formation to simulate strong lensing. That approach naturally yields a realistic distribution of halo and galaxy properties (to the extent that the true distribution of galaxy properties and their relation to halo properties is encoded in the semi-analytic model), at least for a fixed cosmological model, and it will undoubtedly be useful for modeling the strong lensing population in large, future surveys. We elect, however, to take a more controlled approach here, using realistic but discrete values of the density profile parameters, so that we can disentangle the different types of lensing selection biases, and understand the physics of each one, before recombining them to determine the net effects. Our basic approach is to generate realistic galaxy models with various values of the relevant parameters, and to investigate how the lensing cross-section $\\sigma(\\vec{x})$ depends on model parameters $\\vec{x}$ (where $\\vec{x}$ might include mass, shape, concentration, and other parameters, some of which may be correlated). The reason for focusing on the cross-section is that it represents the weighting factor that transforms the intrinsic joint parameter distribution $p(\\vec{x})$ into the distribution $p_\\mathrm{SL}(\\vec{x})$ among observed strong lens systems. Note that we focus on selection biases related to \\emph{physical} effects. There may also be \\emph{observational} selection biases \\citep[e.g., ][]{1991ApJ...379..517K}, but they are specific to a given survey and are not something that we can address in a general way. This fact is one reason for our focus on point-source lensing: extended source lensing is even more sensitive to observational selection effects than quasar lensing. For example, the effect of a finite aperture is no longer straightforward for an extended source, and detection of an extended strong lens system depends on factors such as the size of the point-spread function which varies spatially and temporally in any given survey. Fortunately, with large numbers of point-source lens systems anticipated in future surveys [e.g., $\\sim 10^{5}$ with SKA \\citep{2004NewAR..48.1085K}] this investigation of physical selection biases in point-source lens systems should be highly useful. We emphasize that our purpose is only to determine the mapping function $\\sigma(\\vec{x})$ due to physical differences between galaxies, and not to predict the observed distribution of properties $p_\\mathrm{SL}(\\vec{x})$ in some particular lensing survey, for which knowledge of the intrinsic parameter distribution $p(\\vec{x})$ and of any observational selection effects are both necessary. In this paper, we present a systematic investigation of physical strong lensing selection biases that both unifies and extends previous work. We begin in Section~\\ref{S:simulations} with a brief review of the simulation pipeline developed in Paper~I, including the density profiles and shapes of our galaxy models, our computational lensing methods, and the basic lensing properties of our galaxies. In Section~\\ref{S:orientationshape}, we investigate selection biases related to galaxy orientation and shape. In Section~\\ref{S:innerslope}, we consider selection biases related to the inner slope of the dark matter component of the density profile (for cusped models). In Section~\\ref{S:concentration}, we allow the dark matter concentration to vary as well. In Section~\\ref{S:sersic}, we turn to deprojected S\\'ersic density profiles and examine selection biases with both the stellar and dark matter parameters. Section~\\ref{S:robustness} includes an exploration of the ranges of lens and source redshifts for which our conclusions are applicable. Finally, in Section~\\ref{S:conclusions}, we summarise our main conclusions and discuss their implications for past and future strong lensing analyses. ", "conclusions": "\\label{S:robustness} Finally, to ascertain how much our conclusions about selection biases in the previous sections depend on our chosen fiducial lens and source redshifts, we now try varying the redshifts in several ways. If we vary both the lens and the source redshifts in a way that preserves $\\Sigma_c$, then the physical scale within the galaxy that is measured by strong lensing is the same, but it will lead to different image separation distributions, scaled by the angular diameter distance $D_L$ of the lens galaxy. This change may lead to {\\em observational} selection effects that depend sensitively on the survey image search strategy (resolution and maximum angular separation), but no physical selection effects, so we do not attempt to model this in any general way. Now consider fixing the lens redshift (and therefore the conversion from physical to angular scale), but shifting the source redshift to rescale the critical surface mass density $\\Sigma_c$. This shift will change the Einstein radius $\\Rein$ and therefore the physical scale of the density profile that is probed by strong lensing. Thus, this change may result in different \\emph{physical} selection biases. If we fix the lens redshift to our fiducial $z_L=0.3$ but move the source redshift considerably lower, from the fiducial $z_S=2.0$ to $z_S'=0.6$, then the critical surface density increases by $\\sim 70$ per cent. This change means that the strong lensing is due to smaller scales, where the average projected surface density is higher, and that the lensing cross-section for this density profile will be lower. Because of the importance of smaller scales, the projected dark matter fraction $\\fpdm$ will decrease. As a test, we recompute \\Rein\\ and the lensing cross-sections for both mass scales as a function of the dark matter halo inner slope $\\gdm$ % to see the degree to which the selection bias we identified with our fiducial lens and source redshifts persist. We also increase the resolution of the surface density maps used by \\gravlens\\ for these calculations by a factor of two, while maintaining the box size, to ensure that the critical curves can be properly resolved. We also consider a value of $\\Sigma_c$ that is 40 per cent lower than our fiducial value. While this low $\\Sigma_c$ does not correspond to any sensible value of $z_S$ for our fiducial $z_L$, we assume a higher $z_L\\sim 0.6$ and $z_S\\sim 3$. Because of the different $z_L$, all angular scales output by our pipeline must be rescaled by $D_A(z=0.3)/D_A(z=0.6)=0.66$, i.e., the Einstein radii must be rescaled by $0.66$ and the cross-sections by $0.66^2$. However, the relevant quantities for selection bias associated with true variation of galaxy density profiles, e.g. $\\fpdm$, are preserved. Since we expect increased Einstein radii and cross-sections, we preserve the resolution but increase the box size by a factor of $2$. Table~\\ref{T:varyz} shows what happens to relevant quantities for lensing when we vary $\\Sigma_c$ to values 70 per cent higher and 40 per cent lower. We consider only the unbiased cross-sections for the spherical shape model. \\begin{table*} \\caption{\\label{T:varyz}Summary of the lensing results for the cusped, spherical model when varying $\\Sigma_c$ so that different physical scales are probed. The factor of $0.66$ in \\Rein\\ for the lower $\\Sigma_c$ cases is shown separately since it is not indicative of the physical scale within the galaxy.} \\begin{tabular}{l|r|r|r} \\hline \\hline Quantity & Fiducial $\\Sigma_c$ & Higher $\\Sigma_c$ & Lower $\\Sigma_c$ \\\\ \\hline \\multicolumn{4}{c}{Lower mass scale} \\\\ $\\Rein(\\gdm=0.5)$ & $0.58$ & $0.36$ & $0.79\\times 0.66=0.52$ \\\\ $\\Rein(\\gdm=1.0)$ (arcsec) & $0.68$ & $0.41$ & $0.95\\times 0.66=0.63$ \\\\ $\\Rein(\\gdm=1.5)$ & $1.04$ & $0.61$ & $1.46\\times 0.66=0.97$ \\\\ $\\fpdm(R'<\\Rein,\\gdm=0.5)$ & $0.13$ & $0.08$ & $0.17$ \\\\ $\\fpdm(R'<\\Rein,\\gdm=1.0)$ & $0.28$ & $0.20$ & $0.37$ \\\\ $\\fpdm(R'<\\Rein,\\gdm=1.5)$ & $0.60$ & $0.51$ & $0.67$ \\\\ $\\st(\\gdm=0.5)/\\st(\\gdm=1.0)$ & $0.79$ & $0.79$ & $0.79$ \\\\ $\\st(\\gdm=1.5)/\\st(\\gdm=1.0)$ & $2.65$ & $2.96$ & $2.51$ \\\\ \\hline \\multicolumn{4}{c}{Higher mass scale} \\\\ $\\Rein(\\gdm=0.5)$ & $1.18$ & $0.63$ & $1.73\\times 0.66=1.14$ \\\\ $\\Rein(\\gdm=1.0)$ (arcsec) & $1.75$ & $0.88$ & $2.75\\times 0.66=1.82$ \\\\ $\\Rein(\\gdm=1.5)$ & $4.44$ & $2.33$ & $6.75\\times 0.66=4.46$ \\\\ $\\fpdm(R'<\\Rein,\\gdm=0.5)$ & $0.22$ & $0.13$ & $0.30$ \\\\ $\\fpdm(R'<\\Rein,\\gdm=1.0)$ & $0.51$ & $0.36$ & $0.63$ \\\\ $\\fpdm(R'<\\Rein,\\gdm=1.5)$ & $0.87$ & $0.80$ & $0.91$ \\\\ $\\st(\\gdm=0.5)/\\st(\\gdm=1.0)$ & $0.52$ & $0.51$ & $0.52$ \\\\ $\\st(\\gdm=1.5)/\\st(\\gdm=1.0)$ & $9.19$ & $12.9$ & $8.04$ \\\\ \\hline \\hline \\end{tabular} \\end{table*} First, we can see that when we increase (decrease) $\\Sigma_c$, we decrease (increase) \\Rein\\ and \\fpdm, as expected. The change in $\\Rein$ can be significant, up to a factor of $\\sim 2$ for the changes in $\\Sigma_c$ that we have tested. We find that the selection bias with $\\gdm$, represented in the last two rows of this table, is nearly independent of $\\Sigma_c$ for $\\gdm$ below the fiducial value of $1$, and for higher $\\gdm$ it is a slightly increasing function of $\\Sigma_c$. The sign of this change is surprising, in the sense that if \\fpdm\\ is lower than we might naively expect the parameters of the dark matter distribution to be less significant. However, we can explain this change by the fact that the cross-section \\st\\ is determined by the outer (source-plane) caustic, which relates to the inner (lens-plane) critical curve. This in turn is determined by the slope of the density profile on small scales, so if the lensing properties are determined by smaller (larger) physical scales, as when we increase (decrease) $\\Sigma_c$, then \\gdm\\ becomes more (less) important and selection bias becomes stronger (weaker). However, it is encouraging that the selection bias only changes by $+14$ and $-4$ per cent ($+41$ and $-13$ per cent) for changes in $\\Sigma_c$ of $+70$ and $-40$ per cent for the lower (higher) mass model. This result suggests that within the range of $\\Sigma_c$ that we have considered in this subsection, the conclusions of our paper regarding physical selection bias are not strongly dependent on our choice of fiducial lens and source redshifts. (The same is not true for observational selection biases, which may be strongly affected by changing $z_L$ and thus the characteristic angular scale of the system.) On Fig.~\\ref{F:sigmacrit}, we have indicated the lens and source redshift combinations for which our conclusions are robust according to the tests in this subsection. \\begin{figure} \\begin{center} \\includegraphics[width=\\columnwidth,angle=0]{scinvcontour_shrink.ps} \\caption{\\label{F:sigmacrit}Critical surface density, plotted as $\\log{[\\Sigma_c/(M_{\\odot}/\\mathpc^2)]}$, as a function of lens and source redshift. For $z_S$2~10$^{4}$\\,cm$^{-3}$), high electron temperature (2\\,10$^{4}$\\,K), and low ionisation level (0.03 - 0.3) which varies considerably depending on the target. Indeed, in all cases saturation of the [\\ion{S}{2}] doublet is reached in at least some of the datapoints, and so in this region close to the launch point of the jet the electron density reported is often a lower limit. This is different to findings further along the jet which are a factor of ten lower (\\citealp{Bacciotti99}; \\citealp{Podio06}). In the case where previous studies have been carried out, we find good agreement for values reported close to the star. In the case of TH\\,28 (which presents the best dataset), possible shock signatures are present, thus providing a observational indications that shocks can contribute to heating the jet close to the source. We determine the mass and angular momentum outflow rates for the jets close to their base. Estimates we determined for the mass and angular momentum outflow rates, both of which are fundamental parameters in constraining models of accretion/ejection structures, particularily if the parameters can be determined close to the jet footpoint. Values for a single jet lobe are in the range 4.0~10$^{-9}$ to 6.7~10$^{-8}$\\,M$_\\sun$\\,yr$^{-1}$ and 1.1~10$^{-6}$ to 1.3~10$^{-5}$\\,M$_\\sun$\\,yr$^{-1}$\\,AU\\,km\\,s$^{-1}$. Again we find good agreement with the literature in cases where values were reported close to the star. Mass flow ratios were found to be $\\dot M_{jet}/\\dot M_{acc}$\\,$\\sim$\\,0.01 - 0.07 where accretion rates were available in the literature (i.e. for 3 of 5 targets). This is in the range predicted by accretion-ejection models (\\citealp{Konigl00}; \\citealp{Casse00}; \\citealp{Shu00}). Although we have examined a region of the jet at about 50 - 80\\,AU from the source corresponding to the collimation zone, we note that the region where the jet is formed and launched is believed to be on scales of less than 1\\,AU, a region which is currently out of the reach of present instrumentation and often obscured by infalling matter. We await near-infrared interferometry as an opportunity to observing this zone. \\vspace {0.2in} {\\bf Acknowledgements} \\vspace {0.1in} \\newline The present work was supported in part by the European Community's Marie Curie Actions - Human Resource and Mobility within the JETSET (Jet Simulations, Experiments and Theory) network, under contract MRTN-CT-2004-005592." }, "0808/0808.2612_arXiv.txt": { "abstract": "An analytical method is presented for treating the problem of a uniformly rotating, self-gravitating ring without a central body in Newtonian gravity. The method is based on an expansion about the thin ring limit, where the cross-section of the ring tends to a circle. The iterative scheme developed here is applied to homogeneous rings up to the 20th order and to polytropes with the index $n=1$ up to the third order. For other polytropic indices no analytic solutions are obtainable, but one can apply the method numerically. However, it is possible to derive a simple formula relating mass to the integrated pressure to leading order without specifying the equation of state. Our results are compared with those generated by highly accurate numerical methods to test their accuracy. ", "introduction": "The problem of the self-gravitating ring captured the interest of such renowned scientists as \\citet{Kowalewsky85}, \\citet{Poincare85b,Poincare85c,Poincare85d} and \\citet{Dyson92, Dyson93}. Each of them tackled the problem of an axially symmetric, homogeneous ring in equilibrium by expanding it about the thin ring limit. In particular, Dyson provided a solution to fourth order in the parameter $\\sigma=a/b$, where $a$ provides a measure for the radius of the cross-section of the ring and $b$ the distance of the cross-section's centre of mass from the axis of rotation. An important step toward understanding rings with other equations of state was taken by \\citet{Ostriker64,Ostriker64b,Ostriker65}, who studied polytropic rings to first order in $\\sigma$ and found a complete solution to this order for an isothermal limit. \\par First numerical results for homogeneous rings were given by \\citet{Wong74}, who was not able to clarify the transition to spheroidal bodies that \\citet{Bardeen71} had supposed would exist. \\citet{ES81} and \\citet{EH85} developed improved methods with which they were able to study the connection to the Maclaurin spheroids. Returning to the problem significantly later, \\citet*{AKM03} achieved near-machine accuracy, which allowed them to study bifurcation sequences in detail and correct erroneous results. It was also possible to extend the problem to non-homogeneous rings and even to the framework of General Relativity \\citep*{Hachisu86,AKM03c,FHA05}. \\par Through the use of computer algebra, we extend Dyson's basic idea and determine the solution to the problem of the homogeneous ring up to the order $\\sigma^{20}$. We also present an iterative method for performing a similar expansion about the thin ring limit for arbitrary equations of state and general results are derived, confirming and generalizing work that had already been published by \\citet{Ostriker64b}. The application to polytropes is considered and ordinary differential equations (ODEs) are found that allow for the determination of the mass density. A closed-form solution can only be found if the value of the polytropic index is $n=1$, and such rings are considered to the order $\\sigma^3$. For other polytropic indices, the ODEs are solved numerically so that results from the approximate scheme can be compared to highly accurate numerical results for a variety of equations of state. The numerical solutions considered here are taken from a multi-domain spectral program, much like the one described in \\citet*{AKM03b}, but tailored to Newtonian bodies with toroidal topologies (see \\citealt{AP05} for more information). The solutions obtained by these numerical methods are extremely accurate and thus provide us with a means of testing the accuracy of the approximate method. ", "conclusions": "In their work on homogeneous rings, Poincar\\'e and Kowalewsky, whose results disagreed to first order, both had made mistakes as Dyson has shown. His result to fourth order is also erroneous as we point out in the appendix. It thus seems particularly worthwhile to test the correctness of the solutions presented here. For one thing, we ensured that the transition condition \\begin{align} \\nabla U_\\text{in}|_\\text{s}=\\nabla U_\\text{out}|_\\text{s} \\end{align} is fulfilled up to the appropriate order in $\\sigma$. Furthermore we tested that the virial theorem \\eqref{VI} is fulfilled for each order in $\\sigma$. \\par Please note that one has to be careful in interpreting the results for the thin ring limit. For example, one might think that the squared angular velocity vanishes like $\\sigma^2\\ln\\sigma$. This is true for the dimensionless quantity $\\Omega^2/G\\mu_\\text c$, but need not be true for the squared angular velocity itself. If we fix the `size' $b$ and the mass $M$ of the ring in that limit, then the cross-section shrinks to a point ($a=\\sigma b$). With \\eqref{mass_h00} we can conclude that $\\mu_\\text c\\propto\\sigma^{-2}$ and therefore $\\Omega^2\\propto\\ln\\sigma$, which means that $\\Omega^2$ and hence the velocity of a fluid element tend to infinity. \\par Relativistic rings, including the thin ring limit, were studied in \\citet{AKM03c}, \\citet{Ansorgetal04} and \\citet{FHA05}. From the perspective of General Relativity, the Newtonian theory constitutes a good approximation when certain conditions are fulfilled. For one thing, typical velocities must be small compared to the speed of light $c$ and for another $|U|\\ll c^2$ must hold. We just saw, however, that for rings of finite extent and mass, the velocities grow unboundedly in the thin ring limit. The same holds for $U_\\text s$ as well, see \\eqref{Us_general}. This means that the Newtonian theory of gravity is not appropriate to describe this subtle limit itself, since one cannot expect it to be a good approximation to General Relativity. It is remarkable that the approximation about the point $\\sigma=0$ is nevertheless so successful." }, "0808/0808.2338_arXiv.txt": { "abstract": "Axisymmetric magnetorotational instability (MRI) in viscous accretion disks is investigated by linear analysis and two-dimensional nonlinear simulations. The linear growth of the viscous MRI is characterized by the Reynolds number defined as $R_{\\rm MRI} \\equiv v_A^2/\\nu\\Omega $, where $v_A$ is the Alfv{\\'e}n velocity, $\\nu$ is the kinematic viscosity, and $\\Omega$ is the angular velocity of the disk. Although the linear growth rate is suppressed considerably as the Reynolds number decreases, the nonlinear behavior is found to be almost independent of $R_{\\rm MRI}$. At the nonlinear evolutionary stage, a two-channel flow continues growing and the Maxwell stress increases until the end of calculations even though the Reynolds number is much smaller than unity. A large portion of the injected energy to the system is converted to the magnetic energy. The gain rate of the thermal energy, on the other hand, is found to be much larger than the viscous heating rate. Nonlinear behavior of the MRI in the viscous regime and its difference from that in the highly resistive regime can be explained schematically by using the characteristics of the linear dispersion relation. Applying our results to the case with both the viscosity and resistivity, it is anticipated that the critical value of the Lundquist number $S_{\\rm MRI} \\equiv v_A^2/\\eta\\Omega$ for active turbulence depends on the magnetic Prandtl number $S_{{\\rm MRI},c} \\propto Pm^{1/2}$ in the regime of $Pm \\gg 1$ and remains constant when $Pm \\ll 1$, where $Pm \\equiv S_{\\rm MRI}/R_{\\rm MRI} = \\nu/\\eta$ and $\\eta$ is the magnetic diffusivity. ", "introduction": "Magnetohydrodynamic (MHD) turbulence is the most promising candidate for angular momentum transport in astrophysical disk systems. Nonlinear behaviors of the magnetorotational instability (MRI) are actively investigated as a driving mechanism of MHD turbulence over the last decades (Balbus \\& Hawley 1991, 1998). The central issue in MRI research is its nonlinear properties, in particular, the saturation amplitude of the instability. Although the key processes governing the nonlinear saturation are scoped by global and local numerical studies, it is not fully explained yet (Hawley \\& Balbus 1992; Hawley et al. 1995, 1996; Brandenburg et al. 1995; Matsumoto \\& Tajima 1995; Stone et al. 1996; Hawley 2000; Machida et al. 2000; Arlt \\& R\\\"udiger 2001; Balbus 2003). Recently, Lesur \\& Ogilvie (2008) argue that, in the shearing box context with zero-net vertical flux, MHD turbulence is sustained through nonlinear classical dynamo activity once the MRI is operated. It would be necessary and significative to study the saturation process from the microscopic viewpoint of the physical sustaining mechanism for MHD turbulence as they have done. Ohmic dissipation is one of the crucial processes that determine the saturation amplitude of the MRI. Linear growth rate of the MRI can be reduced significantly because of the suppression by ohmic dissipation. Two- and three-dimensional local shearing box simulations (Sano et al. 1998, 2004; Sano \\& Stone 2002) have shown that physical properties of the saturated turbulence depend on the Lundquist number $S_{\\rm MRI } \\equiv v_A^2/\\eta\\Omega $, where $v_A$ is the Alfv\\'en velocity, $\\Omega $ is the angular velocity, and $\\eta $ is the magnetic diffusivity (see also Fleming et al. 2000; Ziegler \\& R\\\"udiger 2001; Liu et al. 2006). Particularly, when $S_{\\rm MRI} \\ll 1$, the size of the saturated stress decreases with the decrease of $S_{\\rm MRI} $ (Sano \\& Stone 2002; Pessah et al. 2007). It is also pointed out that magnetic reconnection plays an important role in the energy dissipation of MRI driven turbulence (Sano \\& Inutsuka 2001). Numerical issues are one of the main reasons why the saturation physics of the MRI is remained to be understood. Fromang \\& Papaloizou (2007) demonstrate the efficiency of angular momentum transport decreases linearly with the grid spacing as the resolution increases. Although it is very difficult to distinguish between the numerical and physical factors working as the saturation mechanism (King et al. 2007; Silvers 2007), current researches of the MRI pay much attention to the numerical factors with the greatest care. Pessah et al. (2007) derive a scaling law of the saturated stress from past wide variety of numerical results by analytically eliminating the numerical factors such as box size and resolutions. More straightforward way for decontaminating the numerical factors is to bring explicitly the physical diffusivities much larger than the numerical one into the computational study. Lesur \\& Longaretti (2007) have performed first systematic study of the MRI in the presence of both viscous and magnetic dissipations, which are larger than the numerical diffusivities. For the cases with nonzero net flux of the vertical field, the transport property in the saturated state depends on the magnetic Prandtl number $Pm \\equiv \\nu/\\eta $, where $\\nu $ is the kinematic viscosity. Fromang et al. (2007) have reported that in zero magnetic flux cases the turbulent activity is an increasing function of the magnetic Prandtl number $Pm$. Linear behaviors of the MRI in the presence of both the viscosity and resistivity are analytically studied by Pessah \\& Chan (2008) comprehensively. The magnetic Prandtl number $Pm$ takes a wide range of values in astrophysical disk systems. In protoplanetary disks surrounding young stellar objects, the magnetic Prandtl number is much smaller than unity because of their low ionization degree (Nakano 1984; Umebayashi \\& Nakano 1988; Sano et al. 2000). In accretion disks of compact X-ray sources and active galactic nuclei, the magnetic Prandtl number ranges from $\\simeq 10^{-3}$ to $10^{3}$ depending on the distance from the central object (Balbus \\& Henri 2008). In collapsar disks which is known as the central engine of gamma-ray bursts (Woosley 1993), the physical state with high magnetic Prandtl number of $Pm \\gtrsim 10^{10}$ is expected to be realized in their evolutionary stage as a result of the large neutrino viscosity (Masada et al. 2007). Therefore, more systematic and deeper study on the MRI in the presence of both the viscosity and resistivity is quite important for understanding the accretion process triggered by the MRI in various disk systems. One important unsettled matter, in these situations, is the role of the kinematic viscosity at the nonlinear stage of the MRI. In general, the viscosity as well as the magnetic resistivity can suppress the growth of the MRI. However the dependence of nonlinear outcome on the Prandtl number indicates that the role of the viscosity in MRI turbulence could be different from that of the resistivity. Then, the main purpose of this paper is to reveal nonlinear features of the MRI in viscous accretion disks. As is described in what follows, linear growth of the MRI is characterized by the Reynolds number $R_{\\rm MRI} \\equiv v_A^2/\\nu\\Omega $ in the viscous fluid, and by the Lundquist number $S_{\\rm MRI} \\equiv v_A^2/\\eta\\Omega $ in the resistive fluid. Focusing on these two non-dimensional parameters, we clarify the difference in nonlinear behaviors of the MRI between the viscous and resistive systems. Our paper is organized as follows. In \\S~2, linear features of the MRI in the viscous fluid are presented. In \\S~3, nonlinear behavior of the MRI is investigated by two-dimensional MHD simulations taking account of the viscous terms. The differences between the effect of the viscosity and resistivity are also clarified in \\S~3. Finally we make an physical explanation for our nonlinear results with the help of the linear dispersion relation. Applying our results to double diffusive systems, we predict a condition for sustaining active MRI turbulence in the presence of both the viscosity and resistivity in \\S~4. ", "conclusions": "\\subsection{Nonlinear Behavior in the Single Diffusive Systems} To give a physical explanation for the nonlinear behavior of the axisymmetric MRI, we focus on the critical wavelength obtained from the linear theory in this section. The diagrams of Figures~\\ref{fig8}a and \\ref{fig8}b indicate the critical and the fastest growing wavelengths of the MRI as a function of the Lundquist number $S_{\\rm MRI}$ and the Reynolds number $R_{\\rm MRI}$, respectively (see also Table~\\ref{table1}). Note that the vertical axes are normalized by $2\\pi (\\eta/\\Omega)^{1/2} $ in Figure~\\ref{fig8}a and by $2\\pi (\\nu/\\Omega)^{1/2}$ in Figure~\\ref{fig8}b. Shaded area denotes the linearly unstable regions for the MRI. Assuming fixed diffusivities, $S_{\\rm MRI}$ and $R_{\\rm MRI}$ increase as the instability grows because they are proportional to the squared Alfv\\'en velocity. Then the horizontal axis in Figure~\\ref{fig8} can be regarded as the time direction in terms of the evolution of the MRI. First, we consider the resistive case shown by Figure~\\ref{fig8}a. For the case of $S_{\\rm MRI} \\lesssim 1$, the critical wavelength is described as $\\lambda_{\\rm crit} \\simeq \\eta/v_A$. At the nonlinear stage of the two-dimensional MRI, MHD turbulence decays and it saturates only when $S_{\\rm MRI} \\lesssim 1$ (see Figs.~\\ref{fig4}b and \\ref{fig7}). This behavior can be interpreted schematically using the $\\lambda_{\\rm crit}$-$S_{\\rm MRI}$ diagram (Sano \\& Miyama 1999). As the MRI grows and amplifies the magnetic field, the critical wavelength shifts to the shorter length-scale. Then, many smaller scale fluctuations can become unstable. Those structures enhance the efficiency of ohmic dissipation in the turbulent state. In other words, the system evolves toward a more dissipative state, and could be saturated at a critical point around $S_{\\rm MRI} \\simeq 1$, at which the critical wavelength reaches the shortest value and the ohmic dissipation is the most efficient. In this way, MHD turbulence can decay if $S_{\\rm MRI} \\lesssim 1$. When $S_{\\rm MRI} \\gtrsim 1$, on the other hand, the critical wavelength is given by $\\lambda_{\\rm crit} \\simeq v_A/\\Omega $. Two-dimensional calculations of the MRI suggests that the unstable growth cannot saturate if $S_{\\rm MRI} \\gtrsim 1$. The nonlinear behavior in these cases can be explained by Figure~\\ref{fig8}a analogously as follows: At the linear evolutionary stage, the critical wavelength in the radial direction becomes longer due to the exponential growth of the radial field component, while the vertical field grows slower than exponentially. The wavevectors of the unstable modes thus become parallel to the vertical axis. This channel flow mode is the exact solution of the nonlinear MHD equations (Goodman \\& Xu 1994). As the field strength becomes larger, the critical wavelength shifts to the longer one. The dissipation can be much less effective, and thus the channel solution continues to grow without saturation. Next, let us consider the viscous case shown in Figure~\\ref{fig8}b. In the viscous fluid, the critical wavelength is given by $\\lambda_{\\rm crit} \\simeq v_A/\\Omega $ despite the size of the Reynolds number (see Figure~\\ref{fig1}). Even if $R_{\\rm MRI} $ is much smaller than unity, the critical wavelength thus shifts to larger scale as the instability grows. Then the system always evolves toward a less dissipative state and is not saturated. This interpretation is consistent with our numerical results shown in Figure~\\ref{fig7}. These results indicates that the saturation process of the MRI would be changed dramatically at the critical point at which the critical wavelength switches from the decreasing function of the field strength to the increasing one. The differences in the nonlinear behavior of the MRI between the viscous and resistive systems would be originated from whether the critical point exists or not. \\subsection{MRI in the Doubly Diffusive System} In this subsection, we apply the discussion above to the doubly diffusive system that includes both the viscosity and resistivity. The saturation behavior of the axisymmetric MRI can be anticipated by the dependence of the critical wavelength on the field strength derived from the linear theory. With the same framework used in \\S~2, a local axisymmetric dispersion equation of the MRI in the presence of both viscous and ohmic dissipations is given by \\begin{equation} a_4 \\tilde{\\gamma}^4 + a_3 \\tilde{\\gamma}^3 + a_2 \\tilde{\\gamma}^2 + a_1 \\tilde{\\gamma} + a_0 = 0 \\;, \\label{eq20} \\end{equation} where \\begin{eqnarray} a_4 & = & 1\\;, \\ \\ \\ \\ a_3 = 2 \\left( \\frac{1}{R_{\\rm MRI}} + \\frac{1}{S_{\\rm MRI}} \\right) \\tilde{k}_z^2 \\;, \\nonumber \\\\ a_2 & = & \\left( \\frac{1}{R_{\\rm MRI}^2 } + \\frac{1}{S_{\\rm MRI}^2 } + \\frac{4}{R_{\\rm MRI}S_{\\rm MRI} } \\right) \\tilde{k}_z^4+ 2\\tilde{k}_z^2 + \\tilde{\\kappa}^2 \\;, \\nonumber \\\\ a_1 & = & \\left( \\frac{1}{R_{\\rm MRI} } + \\frac{1}{S_{\\rm MRI} } \\right)\\frac{2}{R_{\\rm MRI}S_{\\rm MRI} } \\tilde{k}_z^6 \\nonumber \\\\ && + 2\\left( \\frac{1}{R_{\\rm MRI}} + \\frac{1}{S_{\\rm MRI}}\\right) \\tilde{k}_z^4 + \\frac{2}{S_{\\rm MRI}} \\tilde{k}_z^2\\tilde{\\kappa}^2 \\;, \\nonumber \\\\ a_0 & = & \\frac{1}{R_{\\rm MRI}^2 S_{\\rm MRI}^2} \\tilde{k}_z^8 + \\frac{2}{R_{\\rm MRI}S_{\\rm MRI} }\\tilde{k}_z^6 \\nonumber \\\\ && + \\left( \\frac{1}{S_{\\rm MRI}^2 } \\tilde{\\kappa }^2 + 1 \\right) \\tilde{k}_z^4 + ( \\tilde{\\kappa}^2 - 4) \\tilde{k}_z^2 \\;, \\nonumber \\end{eqnarray} (Menou et al. 2006; Masada et al. 2007; Lesur \\& Longaretti 2007; Pessah \\& Chan 2008). This equation is, as expected, characterized by $R_{\\rm MRI }$ and $S_{\\rm MRI}$. Here we focus on the system with a constant magnetic Prandtl number ($Pm \\equiv S_{\\rm MRI}/R_{\\rm MRI} = \\nu/\\eta $). Figure~\\ref{fig9} demonstrates the critical wavelength of the MRI as a function of $S_{\\rm MRI}$ for various values of $Pm$ obtained by solving the dispersion equation (18). The $S_{\\rm MRI}$-dependence of the critical wavelength varies with the size of $Pm$. When $Pm \\ll 1$, the linear growth of the MRI is independent of the magnetic Prandtl number, and the critical wavelength is almost identical to the pure resistive case ($Pm =0$). However, if the viscosity effects is added sufficiently, then the critical wavelength is enlarged by the suppression due to the viscosity in the middle range of $S_{\\rm MRI}$. For the cases of $Pm \\gg 1$, the critical wavelength around $S_{\\rm MRI} \\simeq 1$ is given by $\\tilde{k}^{-1}_{\\rm crit} \\propto (S_{\\rm MRI}R_{\\rm MRI})^{-1/3} \\propto (S_{\\rm MRI}^2 / Pm)^{-1/3}$ (Pessah \\& Chan 2008). Although the critical wavelength shifts to longer length-scales as the magnetic Prandtl number increases, the critical wavelength has a minimum value for all the cases. This implies that the MRI turbulence could be suppressed if the Lundquist number is less than a critical value. This diagram suggests that the critical Lundquist number $S_{{\\rm MRI},c}$ depends on $Pm$. The critical Lundquist number $S_{{\\rm MRI},c}$ is plotted as a function of $Pm$ in Figure~\\ref{fig10}a. In the regime of $Pm \\gg 1$, it is proportional to the square root of the magnetic Prandtl number, that is $S_{{\\rm crit},c} \\propto Pm^{1/2}$. In contrast, it remains to be constant in the range $Pm \\ll 1$. Since the critical Reynolds number is given by $R_{{\\rm MRI},c} = S_{{\\rm MRI},c}/Pm$, we can also obtain the relation between $Pm$ and $R_{{\\rm MRI},c}$, and which is depicted in Figure~\\ref{fig10}b. Nonlinear growth of the MRI can be expected in the parameter region above this critical curve. The magnetic Prandtl number is proportional to $R_{{\\rm MRI},c}^{-2}$ in the regime of $Pm \\gg 1$ and $Pm \\propto R_{{\\rm MRI},c}^{-1}$ when $Pm \\ll 1$. This curve is reminiscent of the critical curve for MHD turbulence sketched from nonlinear simulations of MRI (Fromang et al. 2007). The Reynolds number $R_{\\rm MRI}$ in their models are at most a few tens\\footnote{The definition of the Reynolds number $Re$ in Fromang et al. (2007) is different from $R_{\\rm MRI}$ in this paper. The relation between these two is $R_{\\rm MRI} \\approx \\alpha_{M} Re$, where $\\alpha_M$ is the Maxwell stress normalized by the (initial) pressure.}, and the critical magnetic Prandtl number is around unity. Thus our prediction is roughly consistent with nonlinear results even quantitatively. Note that the critical curve shown by Figure~\\ref{fig10} is obtained by using only the features of the linear dispersion relation of MRI. This implies that the linear growth of the MRI is very important even at the nonlinear saturated phase to sustain MHD turbulence. The discussion in this paper is based only on two-dimensional simulations of the MRI. For the understanding the saturation mechanism of the MRI, it is quite important to perform the systematic three-dimensional analysis of the MRI in the presence of multiple diffusivities. Furthermore, the assumption of the local shearing box could affect the nonlinear evolution of the viscous MRI. The necessary ingredients of the unstable growth of the MRI are the velocity shear and the magnetic field. In the numerical setting of the local shearing box, the velocity profile of the background shear flow cannot disappear by the role of the kinematic viscosity, but is imposed by the boundary conditions. It would be very interesting to use global disk models to investigate the MRI in highly viscous disks. These are our next tasks." }, "0808/0808.3422_arXiv.txt": { "abstract": "We study dynamics of bars in models of disk galaxies embeded in realistic dark matter halos. We find that disk thickness plays an important, if not dominant, role in the evolution and structure of the bars. We also make extensive numerical tests of different $N$-body codes used to study bar dynamics. Models with thick disks typically used in this type of modeling (height-to-length ratio $h_z/R_d=0.2$) produce slowly rotating, and very long, bars. In contrast, more realistic thin disks with the same parameters as in our Galaxy ($h_z/R_d\\approx 0.1$) produce bars with normal length $R_{\\rm bar} \\approx R_d$, which rotate quickly with the ratio of the corotation radius to the bar radius ${\\cal R} =1.2-1.4$ compatible with observations. Bars in these models do not show a tendency to slow down, and may lose as little as 2-3 percent of their angular momentum due to dynamical friction with the dark matter over cosmological time. We attribute the differences between the models to a combined effect of high phase-space density and smaller Jeans mass in the thin disk models, which result in the formation of a dense central bulge. Special attention is paid to numerical effects such as the accuracy of orbital integration, force and mass resolution. Using three $N$-body codes -- Gadget, ART, and Pkdgrav -- we find that numerical effects are very important and, if not carefully treated, may produce incorrect and misleading results. Once the simulations are performed with sufficiently small time-steps and with adequate force and mass resolution, all the codes produce nearly the same results: we do not find any systematic deviations between the results obtained with TREE codes (Gadget and Pkdgrav) and with the Adaptive-Mesh-Refinement (ART) code. ", "introduction": "Barred galaxies represent a large fraction ($\\sim$65\\% ) of all spiral galaxies \\citep[e.g.,][]{Eskridge00,Sheth08}. Bars are ubiquitous. They are found in all types of spirals: in large lenticular galaxies \\citep{Aguerri05}, in normal spirals such as our Galaxy \\citep{Freudenreich98} and M31 \\citep{Lia06,Beaton07}, and in dwarf magellanic-type galaxies \\citep{Valenzuela07}. An isolated stellar disk embeded into a dark matter halo spontaneously forms a stellar bar as a result of the development of global disk instabilities \\citep[e.g.,][Sec. 6.5]{BT}. Bars continue to be closely scrutinized because of their connection with the dark matter halo \\citep[\\eg][]{od03, hbk05, cvk06, lia07}. Because the bars rotate inside massive dark matter halos, they lose some fraction of the angular momentum to their halos and tend to slow down with time \\citep{Tremaine84,Weinberg85}. The formation of bars and associated pseudobulges is often considered as an {\\it alternative} to the hierarchical clustering model \\citep{Kormendy04}. This appears to be incorrect: recent cosmological simulations indicate that the secular bulge formation is a {\\it part (not an alternative) of the hierarchical scenario}. The simulations of the formation of galaxies in the framework of the standard hierarchical cosmological model indicate that bars form routinely in the course of assembly of halos and galaxies inside them \\citep{Mayer08,Ceverino08}. The simulations have a fine resolution of $\\sim 100$\\,pc and include realistic treatment of gas and stellar feedback, which is important for the survival of a bar. Bars form relatively late: well after the last major merger ($z\\approx 1-2$ for normal spiral such as our Milky Way), when a collision of gas rich galaxies brings lots of gas with substantial angular momentum to the central disk galaxy. As the disk accretes the cold gas from the halo, it forms a new generation of stars and gets more massive. At some stage, the disk becomes massive enough to become unstable to bar formation. Once the stellar bar forms, it exists for the rest of the age of the Universe. The cosmological simulations are still in preliminary stages, and it is likely that many results will change as they become more accurate and the treatment of the stellar feedback improves. However, existing cosmological simulations already show us the place and the role of traditional $N$-body simulations of barred galaxies, which start with an unstable stellar disk. It was not clear whether and how this happens in the real Universe. Now the cosmological hydrodynamic simulations tell us that this is somewhat idealistic, but still a reasonable setup compatible with cosmological models. Simulations of bars play an important role for understanding the phenomenon of barred galaxies \\citep[e.g.,][]{Miller78,Sellwood80,lia03,vk03}. Numerical models successfully account for many observed features of real barred galaxies \\citep{Bureau05,Bureau06,Beaton07}. So, it is very important to assess the accuracy of those simulations. Recent disagreements between results of different research groups \\citep{vk03,od03,sd06} prompted us to undertake a careful testing of numerical effects and to compare results obtained with different codes. This type of code testing is routine in cosmological simulations \\citep{Frenk99,Heitmann05,Heitmann07}, but it never has been done before for bar dynamics. Testing and comparison of numerical codes is important for validating different numerical models. It was instrumental for development of precision cosmology. It is our goal to make such tests for $N$-body models of barred galaxies. We use four different popular $N$-body codes: ART \\citep{KKK97}, Gadget-1 and Gadget-2 \\citep{volker01,volker05}, and Pkdgrav \\citep{wsq04}. ART is an Adaptive Mesh Refinement code that reaches high resolution by creating small cubic cells in areas of high density. Gadget and Pkdgrav, on the other hand, are TREE codes that compute forces directly for nearby particles and use a multipole expansion for distant ones. We use the codes to run a series of simulations using the {\\it same initial conditions} for all codes. We also address another issue: the effects of disk thickness on the structure and evolution of bars. Only recently have the simulations started to have enough mass and force resolution to resolve the vertical height of stellar disks. We use different codes to show that the disk height plays an important and somewhat unexpected role. One of the contentious issues in the simulations of bar dynamics is the angular speed and the structure of bars in massive dark matter halos. The amount and the rate at which bars slow down is still under debate. \\citet{ds98,ds00} find in their massive halo models, i.e., those for which the contributions of the disk and the halo to the mass in the central region are comparable, that the bar loses about 40\\% of its initial angular momentum, $L_z$, in $\\sim 10\\ \\gyr$. However, in simulations with much better force resolution and a more realistic cosmological halo setup, \\citet{vk03} and \\citet{cvk06} find a decrease in $L_z$ of only 4--8\\% in $\\sim 6\\ \\gyr$. \\citet{ds98,ds00} also find that bars do not significantly slow down for lower density halos. \\citet{vk03} presented bar models in which stellar disks were embedded in a CDM Milky-Way-type halos with realistic halo concentrations $c \\sim 15$, where $c$ is the ratio of the virial radius to the characteristic radius of the dark matter halo. These simulations were run with the ART code with high spatial resolution of 20-40 pc. The bars in the models were rotating fast for billions of years. \\citet{vk03} argued that slow bars in previous simulations were an artifact of low resolution. \\citet{sd06} used initial conditions of one of the models of \\citet{vk03} and run a series of simulations using their hybrid, polar-grid code. They found, in most cases, a different evolution than that reported in Valenzuela \\& Klypin. In particular, contrary to Valenzuela \\& Klypin's results, they did not find that the bar pattern speed is almost constant for a long period of time. They attribute the differences to the ART refinement scheme. While \\citet{vk03} mention numerical effects (lack of force and mass resolution) as the cause for excessive slowing down of bars in earlier simulations, there was another effect, which was not noticed by \\citet{vk03}: Their disk was rather thin, with a scale-height $h_z$ of only 0.07 of the disk scale-length $R_d$. This should be compared with $h_z \\sim 0.2\\ R_d$ used in most studies of stellar bars \\citep[e.g.,][]{lia02,lia03,m-v06}. Models of \\citet{ds98} have $h_z = 0.1 R_d$, but the resolution of their simulations was grossly insufficient to resolve this scale. \\citet{od03} used $h_z = 0.1 R_d$ for a model, which had very little dark matter in the central disk region: $M_{\\rm dm}/M_{\\rm disk} \\approx 1/4$ inside radius $R=3R_d$. Dependence of bar speed on disk thickness was noted by \\citet{lia00}: thicker disks result in slower bars. The remainder of this paper is organized as follows. In section~\\ref{sec:height} we give a detailed review of available data on disk scale heights and present a simple analytical model for the relation of the disk scale length with the disk scale height. We describe our numerical models in Section~\\ref{sec:Models}. In Section~\\ref{sec:Numerics} we give a brief description of the codes and present analysis of numerical effects. Main results are presented in Section~\\ref{sec:Results}. We summarize our results in Section~\\ref{sec:Discussion}. ", "conclusions": "\\label{sec:Discussion} \\begin{figure} \\includegraphics[width=0.475\\textwidth]{DensityMapxySK.5.ps} \\includegraphics[width=0.465\\textwidth]{FreudP.ps} \\caption[fig:Surface]{The iso-contours of the surface stellar density for model \\kgthree~ (top panel) and the Milky Way galaxy (bottom panel). The \\kgthree~ model was re-scaled to have the evolved disk scale length 3~kpc, which is close to the scale length of our Galaxy. The bottom panel shows one of the models from \\citet[][fig. 14, right panel]{Freudenreich98}. The model represents a multi parameter fit to the COBE DIRBE maps of the near-infrared light coming from central regions of our Galaxy. The small circle shows position of the Sun. The \\kgthree~ model reproduces the length and the flattening of the Milky Way bar.} \\label{fig:SurfaceMW} \\end{figure} We find that numerical effects -- the mass and force resolution, and the time integration of trajectories -- can significantly alter results of simulations. The time-step of integration must be small enough to allow accurate integration of expected trajectories. Accuracy of energy conservation can give a misleading impression that the simulation is adequate, while it actually makes substantial errors in position of orbits. Our experiments with realistic orbits in models with flat rotation curves indicate that even the best available integration schemes require $\\sim 2000$ time-steps per orbit. This requirement is valid for variable-step schemes with more steps required for a constant-step schemes. Numerical tests with full-scale dynamical models confirm the condition. This condition results in very small time-steps of $dt \\sim 10^4$~yrs. If one uses dimensionless units defined by scaling $G=1$, $M_{\\rm disk}=1$, $R_d =1$, then the dimensionless time step should be $\\tilde dt\\approx 5\\times 10^{-4}$. This should be compared with typically used $\\tilde dt \\approx 0.01$ \\citep[e.g.,][]{lia02}. This time-step would be insufficient for integration of models with central mass concentration presented in this paper \\footnote{Models studied by \\citet{lia02} are less concentrated and do not require a small time-step.}. A constant time-step $dt =5\\times 10^5$~yrs used by \\citet{Widrow08} is too large for models of the Milky Way galaxy studied in their paper. The mass and related with it the force resolution also play important role. As Figure~\\ref{fig:ResCon} shows, a low force resolution produces a less dense central region, which results in a long and a very massive bar. The same model run with better force resolution produces a shorter, less massive, and faster rotating bar. Once the necessary numerical conditions are fulfilled, the codes produce practically the same results. We do not find any systematic deviations between the results obtained with TREE codes (Gadget and Pkdrgav) and with the Adaptive-Mesh-Refinement (ART) code. Disk height is an important parameter, which is often ignored in models of barred galaxies. In the models, which we consider in this paper, the disk height determines the global properties of the bars. Figure~\\ref{fig:Surface} shows surface density maps of the models with different initial disk thickness. Models with thin disks produce short bars with $R_{\\rm bar}\\approx R_d$, which rotate relatively fast: ${\\cal R} =1.2-1.4$ and which show very little decline of the pattern speed. Models with thick disk produce long and slow rotating bar. In order to facilitate the comparison with the our Galaxy, we re-scaled models to have the evolved disk scale length 2.65~kpc and to have the circular velocity at the solar distance 220~km/s. Any $N$-body system has two arbitrary scaling factors, which can be used to scale the system. Having scaled the models to fit the disk scale length, we can compare other parameters of the models. Because the simulations with thin disks produce reasonable models, we use one of the models (\\kgthree) and compare it with the Milky Way. Table~\\ref{tab:MW} gives a list of some parameters. Figure~\\ref{fig:SurfaceMW} compares the surface density maps of the Milky Way \\citep{Freudenreich98} and the \\kgthree~ model. These comparisons show that the model fits the Milky Way reasonably well. We suggest that the disk height is only an indicator of a more fundamental property -- the phase-space density in the central ($R